[go: up one dir, main page]

 
 

Advances in Steel and Composite Steel–Concrete Bridges and Buildings

A special issue of Infrastructures (ISSN 2412-3811).

Deadline for manuscript submissions: closed (31 August 2024) | Viewed by 20824

Special Issue Editor


E-Mail Website
Guest Editor
Mechanics, Sound & Vibration Laboratory, Department of Civil Engineering, College of Engineering, National Taiwan University, Taipei 10617, Taiwan
Interests: behavior of reinforced prestressed concrete and steel structures; bridge engineering; engineering materials; machine learning; finite element method; structural health assessment and monitoring
Special Issues, Collections and Topics in MDPI journals

Special Issue Information

Dear Colleagues,

Construction steel has widely been used worldwide for developing infrastructures, e.g., bridges and buildings, because of its many advantages, including durability, lightweight, high strength, and sustainability. Moreover, combining such advantages with those of concrete, composite steel–concrete structures have increasingly been applied due to a growing demand for new research. In recent years, a great variety of structural members have been developed, including post-tensioned thin-walled steel box-girders, steel–concrete composite decks with shear connectors, and concrete-filled steel tubular and concrete-encased steel members. Particularly, research topics on steel and composite steel–concrete bridges and buildings cover corrosion, fatigue, fire scenarios, limit and ultimate state designs, linear and nonlinear analyses, maintenance, monitoring, post-tensioning applications, progressive collapse, resistance of components, retrofitting and strengthening, seismic, dynamic, and static loadings, stability, etc.

This Special Issue aims to gather new, genuine, and detailed contributions and future perspectives in the aforementioned topics. State-of-the-art papers are also welcome. It is our pleasure to invite you to submit your work and share this call for papers with your colleagues. High-quality manuscripts related to (but not limited to) the following topics in steel and composite steel–concrete bridges and buildings are welcome:

  • Advanced construction technologies;
  • Advanced discrete and finite element modeling;
  • Development of design standards;
  • Development of high-performance material;
  • Laboratory and field investigations;
  • Linear and nonlinear analyses of geometric and material properties;
  • Monitoring techniques/sensor technologies for deterioration conditions;
  • Nondestructive testing methods;
  • Progressive collapse performance;
  • Serviceability issues under seismic, dynamic, and static loadings, fracture, fatigue, fire, corrosion, etc.;
  • Strengthening and repair interventions.

Dr. Marco Bonopera
Guest Editor

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Infrastructures is an international peer-reviewed open access monthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 1800 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • composite structure
  • limit-state behavior
  • linear and nonlinear analysis
  • mechanics
  • numerical modeling
  • post–tensioning
  • service condition
  • stability
  • steel structure
  • structural performance

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (15 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Editorial

Jump to: Research, Other

7 pages, 187 KiB  
Editorial
Advances in Steel and Composite Steel—Concrete Bridges and Buildings
by Marco Bonopera
Infrastructures 2024, 9(10), 169; https://doi.org/10.3390/infrastructures9100169 - 25 Sep 2024
Viewed by 580
Abstract
Construction steel has widely been used worldwide for developing infrastructure, e [...] Full article

Research

Jump to: Editorial, Other

17 pages, 8713 KiB  
Article
Flexural Behavior of Concrete-Filled Steel Tube Beams Composite with Concrete Slab Deck
by Salam Maytham AlObaidi, Mohammed Abbas Mousa, Aqil M. Almusawi, Muhaned A. Shallal and Saif Alzabeebee
Infrastructures 2024, 9(10), 187; https://doi.org/10.3390/infrastructures9100187 - 17 Oct 2024
Viewed by 285
Abstract
Concrete-filled steel tube (CFST) beams have shown their flexural effectiveness in terms of stiffness, strength, and ductility. On the other hand, composite bridge girders demand durable and ductile girders to serve as tension members, while the concrete deck slab resists the compression stresses. [...] Read more.
Concrete-filled steel tube (CFST) beams have shown their flexural effectiveness in terms of stiffness, strength, and ductility. On the other hand, composite bridge girders demand durable and ductile girders to serve as tension members, while the concrete deck slab resists the compression stresses. In this study, six composite CFST beams with concrete slab decks with a span of 170 cm were investigated under a four-point bending test. The main variables of the study were the compressive strength of the concrete deck, the size of CFST beams, and the composite mechanism between the CFST girder and the concrete deck. The results showed that the flexural strength and ductility of the composite system increased by 20% with increasing concrete compressive strength. The study revealed that the higher-strength concrete slab deck enabled the CFST beam to exhibit improved flexural behavior with reduced deflections and enhanced resistance to cracking. The findings also highlighted the importance of considering the interactions between the steel tube and concrete slab deck in determining the flexural behavior of the composite system revealed by strain distribution along the composite beam profile as determined using the digital image correlation DIC technique, where a 40% increase in the flexural strength was obtained when a channel section was added to the joint of the composite section. Full article
Show Figures

Figure 1

Figure 1
<p>Cross-section details of the specimens.</p>
Full article ">Figure 2
<p>Reinforcement details and casting of concrete deck slab.</p>
Full article ">Figure 3
<p>Test setup of the experimental test.</p>
Full article ">Figure 4
<p>Load–displacement curves of the specimens.</p>
Full article ">Figure 5
<p>Load–slip curves of the tested specimens, measured from the side of the beam.</p>
Full article ">Figure 6
<p>Full-field strain distribution of the tested specimens at ultimate load using DIC.</p>
Full article ">Figure 6 Cont.
<p>Full-field strain distribution of the tested specimens at ultimate load using DIC.</p>
Full article ">Figure 7
<p>Horizontal strain distribution along the section of the tested specimens at ultimate load using DIC.</p>
Full article ">Figure 7 Cont.
<p>Horizontal strain distribution along the section of the tested specimens at ultimate load using DIC.</p>
Full article ">Figure 8
<p>Horizontal strain distribution of discontinued and fully connected composite beams, the red colors refer to compression strains and blue colors indicate the tension strains.</p>
Full article ">Figure 9
<p>Strain evolution with loads at multiple locations of the composite beams.</p>
Full article ">Figure 9 Cont.
<p>Strain evolution with loads at multiple locations of the composite beams.</p>
Full article ">
19 pages, 4067 KiB  
Article
Numerical Investigation of the Axial Load Capacity of Cold-Formed Steel Channel Sections: Effects of Eccentricity, Section Thickness, and Column Length
by Diyari B. Hussein and Ardalan B. Hussein
Infrastructures 2024, 9(9), 142; https://doi.org/10.3390/infrastructures9090142 - 26 Aug 2024
Viewed by 611
Abstract
Cold-formed steel channel (CFSC) sections have gained widespread adoption in building construction due to their advantageous properties, including superior energy efficiency, expedited construction timelines, environmental sustainability, material efficiency, and ease of transportation. This study presents a numerical investigation into the axial compressive behavior [...] Read more.
Cold-formed steel channel (CFSC) sections have gained widespread adoption in building construction due to their advantageous properties, including superior energy efficiency, expedited construction timelines, environmental sustainability, material efficiency, and ease of transportation. This study presents a numerical investigation into the axial compressive behavior of CFSC section columns. A rigorously developed finite element model for CFSC sections was validated against existing experimental data from the literature. Upon validation, the model was employed for an extensive parametric analysis encompassing a dataset of 208 CFSC members. Furthermore, the efficacy of the design methodologies outlined in the AISI Specification and AS/NZS Standard were evaluated by comparing the axial load capacities obtained from the numerically generated data with the results of four previously conducted experimental tests. The findings reveal that the codified design equations, based on nominal compressive resistances determined using the current direct strength method, exhibit a conservative bias. On average, these equations underestimate the actual load capacities of CFSC section columns by approximately 11.5%. Additionally, this investigation explores the influence of eccentricity, cross-sectional dimensions, and the point-of-load application on the axial load capacity of CFSC columns. The results demonstrate that a decrease in section thickness, an increase in column length, and a higher degree of eccentricity significantly reduce the axial capacity of CFSC columns. Full article
Show Figures

Figure 1

Figure 1
<p>Types of Buckling in Thin-Walled Steel Columns.</p>
Full article ">Figure 2
<p>Finite element model parameters, boundary conditions, and cross-section geometry.</p>
Full article ">Figure 3
<p>Engineering and True Stress–Strain Curves.</p>
Full article ">Figure 4
<p>Comparison of load-displacement curves for fixed-ended columns between experimental tests [<a href="#B34-infrastructures-09-00142" class="html-bibr">34</a>] and FEM.</p>
Full article ">Figure 5
<p>Comparison of column buckling shapes: experimental testing [<a href="#B34-infrastructures-09-00142" class="html-bibr">34</a>] vs. finite element analysis.</p>
Full article ">Figure 6
<p>Correlation of axial load capacity with column thickness.</p>
Full article ">Figure 7
<p>The correlation between column length and axial load capacity.</p>
Full article ">Figure 8
<p>Effect of strong-axis eccentricity on column axial load capacity.</p>
Full article ">Figure 9
<p>Effect of weak-axis eccentricity on column axial load capacity.</p>
Full article ">
22 pages, 3664 KiB  
Article
Buckling Instability of Monopiles in Liquefied Soil via Structural Reliability Assessment Framework
by Brian Bachinilla, Milind Siddhpura, Ana Evangelista, Ahmed WA Hammad and Assed N. Haddad
Infrastructures 2024, 9(8), 123; https://doi.org/10.3390/infrastructures9080123 - 26 Jul 2024
Viewed by 976
Abstract
During devastating earthquakes, soil liquefaction has disastrous outcomes on bridge foundations, as mentioned in books and published research. To avoid foundation failure when the surrounding soil is fully liquefied, a bridge’s pile foundation design could be such that the bridge pier is directly [...] Read more.
During devastating earthquakes, soil liquefaction has disastrous outcomes on bridge foundations, as mentioned in books and published research. To avoid foundation failure when the surrounding soil is fully liquefied, a bridge’s pile foundation design could be such that the bridge pier is directly resting on the top of a large-diameter monopile instead of the traditional multiple small-diameter piles. This paper discusses the gap of insufficient studies on large-diameter monopiles to support railway bridges subjected to buckling instability and the lack of simplified tools to quickly assess structural reliability. A framework could quickly assess the structural reliability by formulating a simplified reliability analysis. This study focused on pure buckling with shear deformation and reliability assessment to calculate a monopile’s failure probability in fully liquefied soils. In reliability assessment, with the critical pile length (Lcrit) and the unsupported pile length (Luns), the limit state function g(x) = [Lcrit − Luns] thus forms the basis for assessing the safety and reliability of a structure, indicating the state of success or failure. The Lcrit formulation is accomplished with a differential equation. Here, Luns assumes various depths of liquefied soil. The reliability index’s (β) formulation is achieved through the Hasofer–Lind concept and then double-checked through a normal or Gaussian distribution. A case study was conducted using a high-speed railway bridge model from a published research to demonstrate the application of the proposed methodology. To validate the minimum pile diameter for buckling instability when a fully liquefied soil’s thickness reaches the condition that Lcrit = Luns, this study applies the published research of Bhattacharya and Tokimatsu. The validation results show good agreement for 0.85–0.90 m monopile diameters. With a monopile diameter smaller than 0.85 m, the Lcrit = Luns limit was at lesser depths, while with a monopile diameter larger than 0.90 m, the Lcrit = Luns limit was at deeper depths. A load increase notably affected the large-diameter monopiles because the Lcrit movement required a longer range. In fully liquefied soil, buckling will likely happen in piles with a diameter between 0.50 m and 1.60 m because the calculated probability of failure (Pf) value is nearly one. Conversely, buckling instability will likely not happen in monopiles with a diameter of 1.80–2.20 m because the Pf value is zero. Hence, the outcome of this case study suggests that the reliable monopile minimum diameter is 1.80 m for supporting a high-speed railway bridge. Lastly, this paper analyzed the shear deformation effect on large-diameter monopiles, the result of which was 0.30% of Lcrit. Shear deformation makes minimal contributions to large-diameter monopile buckling. Full article
Show Figures

Figure 1

Figure 1
<p>Bridge pier on a large-diameter monopile [<a href="#B1-infrastructures-09-00123" class="html-bibr">1</a>].</p>
Full article ">Figure 2
<p>Mathematical formulation framework of the current study.</p>
Full article ">Figure 3
<p>Critical pile length and unsupported pile length hypothesis.</p>
Full article ">Figure 4
<p>Shearing deformation on a column element [<a href="#B42-infrastructures-09-00123" class="html-bibr">42</a>].</p>
Full article ">Figure 5
<p>Hasofer-Lind Reliability Index for a linear state function [<a href="#B38-infrastructures-09-00123" class="html-bibr">38</a>].</p>
Full article ">Figure 6
<p>Validation of the current study using Bhattacharya and Tokimatsu’s study [<a href="#B40-infrastructures-09-00123" class="html-bibr">40</a>].</p>
Full article ">Figure 7
<p>Critical pile length comparison of the current study.</p>
Full article ">Figure 8
<p>Probability of failure of the current study.</p>
Full article ">Figure 9
<p>Shear deformation effect of the current study.</p>
Full article ">
26 pages, 23326 KiB  
Article
Fatigue Consideration for Tension Flange over Intermediate Support in Skewed Continuous Steel I-Girder Bridges
by Dariya Tabiatnejad, Seyed Saman Khedmatgozar Dolati, Armin Mehrabi and Todd A. Helwig
Infrastructures 2024, 9(7), 99; https://doi.org/10.3390/infrastructures9070099 - 26 Jun 2024
Viewed by 1135
Abstract
Skewed supports complicate load paths in continuous steel I-girder bridges, causing secondary stresses and differential deformations. For a continuous bridge where tensile stresses are developed in the top flange of the steel girders over the intermediate supports, these effects may exacerbate potential fatigue [...] Read more.
Skewed supports complicate load paths in continuous steel I-girder bridges, causing secondary stresses and differential deformations. For a continuous bridge where tensile stresses are developed in the top flange of the steel girders over the intermediate supports, these effects may exacerbate potential fatigue issues for the top flanges. There is a gap in knowledge regarding the level of stress one can expect at these locations, and the stress level can render the problem either serious or trivial. This paper has been successful in providing this information, which was not available before. The study examines the fatigue performance of the top flange in girders over skewed supports. Results are presented from a detailed investigation consisting of 3D finite element modeling to evaluate 26 skewed bridges in the State of Florida that represent the wide range of geometries found in practice. The analysis focused on stress ranges in the top flanges and axial demands on end cross-frame members under fatigue truck loading. A preliminary analysis helped to select the appropriate element type and support conditions. The maximum factored stress range of 3.63 ksi obtained for the selected group of bridges remains below the 10 ksi fatigue threshold for an AASHTO Category C connection, alleviating the concerns about the fatigue performance of the continuous girder top flange over the intermediate pier. Hence, fatigue is unlikely to be a concern in the flanges at this location. Statistics on computed stress ranges and cross-frame forces that provide an understanding of the expected values and guidance for detailing practices are also presented. A limited comparative refined FE analysis on two different types of end cross-frame to girder connections also provided useful insight into the fatigue sensitivities of the skew connections. Half-Round Bearing Stiffener (HRBS) connections performed better than the customary bent plate connections. The HRBS connection reduces girder flange stress concentration range by at least 18% compared to the bent plate connection. The maximum stress concentration range in bent plate components is significantly higher than in the HRBS connection components. The work documented in this paper is important for understanding the fatigue performance of the cross-frames and girders in support regions in the upcoming 10th edition of the AASHTO Bridge Design Specifications that may include plate stiffeners oriented either normally or skewed to the girder web, or Half-Round Bearing Stiffeners. Full article
Show Figures

Figure 1

Figure 1
<p>A skew angle in a steel I-girder bridge under construction before casting concrete deck.</p>
Full article ">Figure 2
<p>An example of fatigue-prone zone at the connection of the stiffener to the girder web [<a href="#B7-infrastructures-09-00099" class="html-bibr">7</a>].</p>
Full article ">Figure 3
<p>An example of fatigue-prone zone at the connection of the stiffener to the steel I-girder flange [<a href="#B2-infrastructures-09-00099" class="html-bibr">2</a>,<a href="#B8-infrastructures-09-00099" class="html-bibr">8</a>].</p>
Full article ">Figure 4
<p>X-type cross-frame configuration (No scale).</p>
Full article ">Figure 5
<p>K-frame Cross-Frame Configuration (No scale).</p>
Full article ">Figure 6
<p>Full-depth diaphragm member detailing (No scale).</p>
Full article ">Figure 7
<p>End cross-frame location in a steel bridge with 60° skew in Lubbock, TX [<a href="#B17-infrastructures-09-00099" class="html-bibr">17</a>].</p>
Full article ">Figure 8
<p>Skewed end cross-frame bent plate connection [<a href="#B17-infrastructures-09-00099" class="html-bibr">17</a>].</p>
Full article ">Figure 9
<p>Bent plate end cross-frame connection.</p>
Full article ">Figure 10
<p>Small-scale connection test specimens by Quadrato et al. [<a href="#B17-infrastructures-09-00099" class="html-bibr">17</a>].</p>
Full article ">Figure 11
<p>Large-scale test specimen with the HRBS (<b>left</b>) and a single stiffener prepared for welding (<b>right</b>) Quadrato et al. [<a href="#B17-infrastructures-09-00099" class="html-bibr">17</a>].</p>
Full article ">Figure 12
<p>Finite element models of the HRBS connection (<b>left</b>) and an end cross-frame to girders (<b>right</b>) Quadrato et al. [<a href="#B17-infrastructures-09-00099" class="html-bibr">17</a>].</p>
Full article ">Figure 13
<p>3D FEM steel I-girder bridge model.</p>
Full article ">Figure 14
<p>3D FEM steel I-girder bridge model using plate and beam elements.</p>
Full article ">Figure 15
<p>HS–20 Design truck load.</p>
Full article ">Figure 16
<p>Influence surface for the longitudinal force in the top flange element over the skewed support in the first lane of Bridge No. 23.</p>
Full article ">Figure 17
<p>The optimal position of the fatigue truck in the first lane of Bridge No. 23 to maximize stress in the top flange of the steel I–girder over the skewed support.</p>
Full article ">Figure 18
<p>The optimal position of the fatigue truck in the first lane of Bridge No. 23 to minimize stress in the top flange of the steel I–girder over the skewed support.</p>
Full article ">Figure 19
<p>Maximum top-flange Fatigue–I envelope stress range (ksi).</p>
Full article ">Figure 20
<p>Maximum tensile forces in skewed end cross-frame members of Bridge 23 under Fatigue–I envelope loading conditions (measured in ksi).</p>
Full article ">Figure 21
<p>Maximum compressive forces in skewed end cross-frame members of Bridge 23 under Fatigue–I envelope loading conditions (measured in ksi).</p>
Full article ">Figure 22
<p>Tensile axial force distribution in top chord elements for Fatigue–I envelope (kips).</p>
Full article ">Figure 23
<p>Compressive axial force distribution in top chord elements for Fatigue–I envelope (kips).</p>
Full article ">Figure 24
<p>Tensile axial force distribution in diagonal elements for Fatigue–I envelope (kips).</p>
Full article ">Figure 25
<p>Compressive axial force distribution in diagonal elements for Fatigue–I envelope (kips).</p>
Full article ">Figure 26
<p>Tensile axial force distribution in bottom chord elements for Fatigue–I envelope (kips).</p>
Full article ">Figure 27
<p>Compressive axial force distribution in bottom chord elements for Fatigue–I envelope (kips).</p>
Full article ">
22 pages, 7410 KiB  
Article
Bond Stress Behavior of a Steel Reinforcing Bar Embedded in Geopolymer Concrete Incorporating Natural and Recycled Aggregates
by Qasim Shaukat Khan, Haroon Akbar, Asad Ullah Qazi, Syed Minhaj Saleem Kazmi and Muhammad Junaid Munir
Infrastructures 2024, 9(6), 93; https://doi.org/10.3390/infrastructures9060093 - 31 May 2024
Cited by 2 | Viewed by 800
Abstract
The rise in greenhouse gases, particularly carbon dioxide (CO2) emissions, in the atmosphere is one of the major causes of global warming and climate change. The production of ordinary Portland cement (OPC) emits harmful CO2 gases, which contribute to sporadic [...] Read more.
The rise in greenhouse gases, particularly carbon dioxide (CO2) emissions, in the atmosphere is one of the major causes of global warming and climate change. The production of ordinary Portland cement (OPC) emits harmful CO2 gases, which contribute to sporadic heatwaves, rapid melting of glaciers, flash flooding, and food shortages. To address global warming and climate change challenges, this research study explores the use of a cement-less recycled aggregate concrete, a sustainable approach for future constructions. This study uses fly ash, an industrial waste of coal power plants, as a 100% substitute for OPC. Moreover, this research study also uses recycled coarse aggregates (RCAs) as a partial to complete replacement for natural coarse aggregates (NCAs) to preserve natural resources for future generations. In this research investigation, a total of 60 pull-out specimens were prepared to investigate the influence of steel bar diameter (9.5 mm, 12.7 mm, and 19.1 mm), bar embedment length, db (4db and 6db), and percentage replacements of NCA with RCA (25%, 50%, 75%, and 100%) on the bond stress behavior of cement-less RA concrete. The test results exhibited that the bond stress of cement-less RCA concrete decreased by 6% with increasing steel bar diameter. Moreover, the bond stress decreased by 5.5% with increasing bar embedment length. Furthermore, the bond stress decreased by 7.6%, 7%, 8.8%, and 20.4%, respectively, with increasing percentage replacements (25%, 50%, 75%, and 100%) of NCA with RCA. An empirical model was developed correlating the bond strength to the mean compressive strength of cement-less RCA concrete, which matched well with the experimental test results and predictions of the CEB-FIP model for OPC. The CRAC mixes exhibited higher costs but significantly lower embodied CO2 emissions than OPC concrete. Full article
Show Figures

Figure 1

Figure 1
<p>The materials used in GPC are (<b>a</b>) fly ash, (<b>b</b>) sand, (<b>c</b>) NCA, and (<b>d</b>) RCA.</p>
Full article ">Figure 2
<p>Concrete crusher machine.</p>
Full article ">Figure 3
<p>Stages in the preparation of alkaline activator solution used in GPC.</p>
Full article ">Figure 4
<p>Detail of pull-out test specimen.</p>
Full article ">Figure 5
<p>Steel bars enveloped with PVC tubes.</p>
Full article ">Figure 6
<p>Cylindrical molds placed on a mechanical vibrating table before concrete pouring.</p>
Full article ">Figure 7
<p>Demolded samples stored at room temperature.</p>
Full article ">Figure 8
<p>Pull-out test setup.</p>
Full article ">Figure 9
<p>Yielding/bond failure in tested pull-out specimens.</p>
Full article ">Figure 10
<p>Splitting failure in pull-out test specimens.</p>
Full article ">Figure 11
<p>Influence of embedment lengths on bond stress of CRAC mixes.</p>
Full article ">Figure 12
<p>Influence of varying diameters of embedded steel bar in CRAC mixes.</p>
Full article ">Figure 13
<p>Influence of varying percentage replacements of NCA with RCA in the CRAC mix.</p>
Full article ">Figure 14
<p>Stress–slip curves of tested pull-out specimens.</p>
Full article ">Figure 15
<p>Analytical predictions of bond stress with varying compressive strength using existing models and the proposed model [<a href="#B25-infrastructures-09-00093" class="html-bibr">25</a>,<a href="#B48-infrastructures-09-00093" class="html-bibr">48</a>,<a href="#B49-infrastructures-09-00093" class="html-bibr">49</a>,<a href="#B51-infrastructures-09-00093" class="html-bibr">51</a>,<a href="#B52-infrastructures-09-00093" class="html-bibr">52</a>].</p>
Full article ">
29 pages, 10022 KiB  
Article
The Influence of Soil Deformability on the Seismic Response of 3D Mixed R/C–Steel Buildings
by Paraskevi K. Askouni
Infrastructures 2024, 9(5), 80; https://doi.org/10.3390/infrastructures9050080 - 4 May 2024
Viewed by 1172
Abstract
Following effective seismic codes, common buildings are considered to be made of the same material throughout the story distribution and based on an ideal rigid soil. However, in daily construction practice, there are often cases of buildings formed by a bottom part constructed [...] Read more.
Following effective seismic codes, common buildings are considered to be made of the same material throughout the story distribution and based on an ideal rigid soil. However, in daily construction practice, there are often cases of buildings formed by a bottom part constructed with reinforced concrete (r/c) and a higher steel part, despite this construction type not being recognized by code assumptions. In addition, soil deformability, commonly referred to as the Soil–Structure Interaction (SSI), is widely found to affect the earthquake response of typical residence structures, apart from special structures, though it is not included in the normative design procedure. This work studies the seismic response of in-height mixed 3D models, considering the effect of sustaining deformable ground compared to the common rigid soil hypothesis, which has not been clarified so far in the literature. Two types of soft soil, as well as the rigid soil assumption, acting as a reference point, are considered, while two limit interconnections between the steel part on the concrete part are included in the group analysis. The possible influence of the seismic orientation angle is explored in the analysis set. Selected numerical results of the dynamic nonlinear analyses under strong near-fault ground excitations were plotted through dimensionless parameters to facilitate an objective comparative discussion. The effect of SSI on the nonlinear performance of three-dimensional mixed models is identified, which serves as the primary contribution of this work, making it unique among the numerous research works available globally and pointing to findings that are useful for the enhancement of the seismic rules regarding the design and analysis of code-neglected mixed buildings. Full article
Show Figures

Figure 1

Figure 1
<p>The considered mixed frames (<b>a</b>) RC1-ST1, (<b>b</b>) RC2-ST1, (<b>c</b>) RC3-ST1, (<b>d</b>) RC3-ST2, and (<b>e</b>) RC4-ST2, where gray refers to reinforced concrete as structural material and blue refers to structural steel, with a notation of the global axes system.</p>
Full article ">Figure 2
<p>Detail of the 3D orientation of steel elements of the steel stories.</p>
Full article ">Figure 3
<p>Discrete SSI model formed by rigid elements.</p>
Full article ">Figure 4
<p>A Comparison of the considered earthquake spectra in the two horizontal directions (labeled as “1” or “2” for each earthquake) to the code design spectra considering zone ground acceleration of 0.36·g for soil C (labeled as “EC8-Soil C”) and the respective one for soil D (labeled as “EC8-Soil D”).</p>
Full article ">Figure 5
<p>Comparison of the IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 1st story, RC1-ST1 building.</p>
Full article ">Figure 6
<p>Comparison of the IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 2nd story, RC1-ST1 building.</p>
Full article ">Figure 7
<p>Comparison of the base shear ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC1-ST1 building.</p>
Full article ">Figure 8
<p>Comparison of the base moment ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC1-ST1 building.</p>
Full article ">Figure 9
<p>Comparison of the ratios Vb(rel)/Vb(uni) and Mb(rel)/Mb(uni) at the X- and Y-axes, RC1-ST1 building, for “r.s.”, “C”, and “D” soil assumptions.</p>
Full article ">Figure 10
<p>Comparison of IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 1st story, RC2-ST1 building.</p>
Full article ">Figure 11
<p>Comparison of IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 2nd story, RC2-ST1 building.</p>
Full article ">Figure 12
<p>Comparison of the IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 3rd story, RC2-ST1 building.</p>
Full article ">Figure 13
<p>Comparison of the base shear ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC2-ST1 building.</p>
Full article ">Figure 14
<p>Comparison of the base moment ratio along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC2-ST1 building.</p>
Full article ">Figure 15
<p>Comparison of the ratios Vb(rel)/Vb(uni) and Mb(rel)/Mb(uni) on the X- and Y-axes at the RC2-ST1 building for “r.s.”, “C”, and “D” soil assumptions.</p>
Full article ">Figure 16
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 1st story, RC3-ST1 building.</p>
Full article ">Figure 17
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 2nd story, RC3-ST1 building.</p>
Full article ">Figure 18
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 4th story, RC3-ST1 building.</p>
Full article ">Figure 19
<p>Comparison of the base shear ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC3-ST1 building.</p>
Full article ">Figure 20
<p>Comparison of the base moment ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC3-ST1 building.</p>
Full article ">Figure 21
<p>Comparison of the ratios Vb(rel)/Vb(uni) and Mb(rel)/Mb(uni) on the X- and Y-axes at the 4 RC3-ST1 building for “r.s.”, “C”, and “D” soil assumptions.</p>
Full article ">Figure 22
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 1st story, RC3-ST2 building.</p>
Full article ">Figure 23
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 4th story, RC3-ST2 building.</p>
Full article ">Figure 24
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 5th story, RC3-ST2 building.</p>
Full article ">Figure 25
<p>Comparison of the base shear ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC3-ST2 building.</p>
Full article ">Figure 26
<p>Comparison of the base moment ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y axis, RC3-ST2 building.</p>
Full article ">Figure 27
<p>Comparison of the ratios Vb(rel)/Vb(uni) and Mb(rel)/Mb(uni) on the X- and Y-axes at the RC3-ST2 building for “r.s.”, “C”, and “D” soil assumptions.</p>
Full article ">Figure 28
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 1st story, RC4-ST2 building.</p>
Full article ">Figure 29
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 5th story, RC4-ST2 building.</p>
Full article ">Figure 30
<p>IDR along the (<b>a</b>) X-axis and (<b>b</b>) Y-axis at the 6th story, RC4-ST2 building.</p>
Full article ">Figure 31
<p>Comparison of the base shear ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC4-ST2 building.</p>
Full article ">Figure 32
<p>Comparison of the base moment ratio on the (<b>a</b>) X-axis and (<b>b</b>) Y-axis, RC4-ST2 building.</p>
Full article ">Figure 33
<p>Comparison of the Vb(rel)/Vb(uni) and Mb(rel)/Mb(uni) on the X- and Y-axes at the RC4-ST2 building for “r.s.”, “C”, and “D” soil assumptions.</p>
Full article ">Figure 34
<p>Hinge formation for the mixed model RC2-ST1 for (<b>a</b>) rigid soil, release connection, Imperial Valley earthquake with 0°, (<b>b</b>) soil C, release connection, Cape Mendocino earthquake with 90°, and (<b>c</b>) soil D, uniform connection, Cape Mendocino earthquake with 90°.</p>
Full article ">Figure 35
<p>Hinge formation for the mixed model RC3-ST1 for (<b>a</b>) rigid soil, uniform connection, San Fernando earthquake with 0°, (<b>b</b>) soil C, uniform connection, San Fernando earthquake with 90°, and (<b>c</b>) soil D, release connection, San Fernando earthquake with 0°.</p>
Full article ">
33 pages, 8700 KiB  
Article
Enhancing Flexural Strength of RC Beams with Different Steel–Glass Fiber-Reinforced Polymer Composite Laminate Configurations: Experimental and Analytical Approach
by Arash K. Pour, Mehrdad Karami and Moses Karakouzian
Infrastructures 2024, 9(4), 73; https://doi.org/10.3390/infrastructures9040073 - 12 Apr 2024
Cited by 1 | Viewed by 1620 | Correction
Abstract
This study intended to measure the efficiency of different strengthening techniques to advance the flexural characteristics of reinforced concrete (RC) beams using glass fiber-reinforced polymer (GFRP) laminates, including externally bonded reinforcement (EBR), externally bonded reinforcement on grooves (EBROG), externally bonded reinforcement in grooves [...] Read more.
This study intended to measure the efficiency of different strengthening techniques to advance the flexural characteristics of reinforced concrete (RC) beams using glass fiber-reinforced polymer (GFRP) laminates, including externally bonded reinforcement (EBR), externally bonded reinforcement on grooves (EBROG), externally bonded reinforcement in grooves (EBRIG), and the near-surface mounted (NSM) system. A new NSM technique was also established using an anchorage rebar. Then, the effect of the NSM method with and without externally strengthening GFRP laminates was studied. Twelve RC beams (150 × 200 × 1500 mm) were manufactured and examined under a bending system. One specimen was designated as the control with no GFRP laminate. To perform the NSM method, both steel and GFRP rebars were used. In the experiments, capability, as well as the deformation and ductileness of specimens, were evaluated, and a comparison was made between the experimental consequences and existing standards. Finally, a new regression was generated to predict the final resistance of RC beams bound with various retrofitting techniques. The findings exhibited that the NSM technique, besides preserving the strengthening materials, could enhance the load-bearing capacity and ductileness of RC beams up to 42.3% more than the EBR, EBROG, and EBRIG performances. Full article
Show Figures

Figure 1

Figure 1
<p>Advantages and disadvantages of strengthening methods.</p>
Full article ">Figure 2
<p>General overview of the current study program.</p>
Full article ">Figure 3
<p>Geometry of the specimens and rebar setup.</p>
Full article ">Figure 4
<p>Test setup.</p>
Full article ">Figure 5
<p>Strengthening with GFRP-EBR.</p>
Full article ">Figure 6
<p>Strengthening with GFRP-EBROG.</p>
Full article ">Figure 7
<p>Strengthening with GFRP-EBRIG.</p>
Full article ">Figure 8
<p>Strengthening with GFRP-UNSM.</p>
Full article ">Figure 9
<p>Strengthening with GFRP-ANSM.</p>
Full article ">Figure 10
<p>Simultaneous influence of both external strengthening rebar and GFRP laminates: (<b>a</b>) UNSM method with GFRP laminates and (<b>b</b>) ANSM method with GFRP laminates.</p>
Full article ">Figure 11
<p>Load–displacement behavior of beams strengthened with various techniques (<b>a</b>) specimens strengthened with GFRP laminates, (<b>b</b>) specimens strengthened with additional rebars and (<b>c</b>) Specimens strengthened with both additional rebar and GFRP laminates.</p>
Full article ">Figure 11 Cont.
<p>Load–displacement behavior of beams strengthened with various techniques (<b>a</b>) specimens strengthened with GFRP laminates, (<b>b</b>) specimens strengthened with additional rebars and (<b>c</b>) Specimens strengthened with both additional rebar and GFRP laminates.</p>
Full article ">Figure 12
<p>Impacts of various strengthening methods on the bending characteristics of RC beams: (<b>a</b>) final displacement and (<b>b</b>) final load resistance.</p>
Full article ">Figure 13
<p>Modes of failure and crack propagation mid-span.</p>
Full article ">Figure 14
<p>Description of the ductility ratio.</p>
Full article ">Figure 15
<p>Influence of various strengthening techniques on ductility ratios of beams.</p>
Full article ">Figure 16
<p>Influence of the various strengthening techniques on the stiffness of the beams.</p>
Full article ">Figure 17
<p>Strain and stress distributions over the cross-section under the balanced state.</p>
Full article ">Figure 18
<p>Strain and stress over the cross-section when the external strengthening rebar (steel or GFRP) fails.</p>
Full article ">Figure 19
<p>Strain and stress distributions over beams caused by compression failure.</p>
Full article ">Figure 20
<p>Experimental load resistance versus forecast formulas for EBR-strengthened specimens. (<b>a</b>) ACI440.2R 17 [<a href="#B58-infrastructures-09-00073" class="html-bibr">58</a>], (<b>b</b>) Deng et al, [<a href="#B39-infrastructures-09-00073" class="html-bibr">39</a>] (<b>c</b>) CNR-DT 200 R1/2013 [<a href="#B64-infrastructures-09-00073" class="html-bibr">64</a>], (<b>d</b>) Said and Wu [<a href="#B65-infrastructures-09-00073" class="html-bibr">65</a>], (<b>e</b>) Lu et al. [<a href="#B66-infrastructures-09-00073" class="html-bibr">66</a>] and (<b>f</b>) Teng et al. [<a href="#B67-infrastructures-09-00073" class="html-bibr">67</a>].</p>
Full article ">Figure 20 Cont.
<p>Experimental load resistance versus forecast formulas for EBR-strengthened specimens. (<b>a</b>) ACI440.2R 17 [<a href="#B58-infrastructures-09-00073" class="html-bibr">58</a>], (<b>b</b>) Deng et al, [<a href="#B39-infrastructures-09-00073" class="html-bibr">39</a>] (<b>c</b>) CNR-DT 200 R1/2013 [<a href="#B64-infrastructures-09-00073" class="html-bibr">64</a>], (<b>d</b>) Said and Wu [<a href="#B65-infrastructures-09-00073" class="html-bibr">65</a>], (<b>e</b>) Lu et al. [<a href="#B66-infrastructures-09-00073" class="html-bibr">66</a>] and (<b>f</b>) Teng et al. [<a href="#B67-infrastructures-09-00073" class="html-bibr">67</a>].</p>
Full article ">Figure 21
<p>Comparison with experimental results and those obtained using Equation (42).</p>
Full article ">Figure 22
<p>Existing errors between experimental and numerical results for beams tested in this research.</p>
Full article ">
22 pages, 6290 KiB  
Article
Joint Behavior of Full-Scale Precast Concrete Pipe Infrastructure: Experimental and Numerical Analysis
by Abdul Basit, Safeer Abbas, Muhammad Mubashir Ajmal, Ubaid Ahmad Mughal, Syed Minhaj Saleem Kazmi and Muhammad Junaid Munir
Infrastructures 2024, 9(4), 69; https://doi.org/10.3390/infrastructures9040069 - 3 Apr 2024
Cited by 2 | Viewed by 1647
Abstract
This study undertakes a comprehensive experimental and numerical analysis of the structural integrity of buried RC sewerage pipes, focusing on the performance of two distinct jointing materials: cement mortar and non-shrinkage grout. Through joint shear tests on full-scale sewer pipes under single point [...] Read more.
This study undertakes a comprehensive experimental and numerical analysis of the structural integrity of buried RC sewerage pipes, focusing on the performance of two distinct jointing materials: cement mortar and non-shrinkage grout. Through joint shear tests on full-scale sewer pipes under single point loading conditions, notable effects on the crown and invert of the joint were observed, highlighting the critical vulnerability of these structures to internal and external pressures. Two materials—cement–sand mortar and non-shrinkage grout—were used in RC pipe joints to experimentally evaluate the joint strength of the sewerage pipes. Among the materials tested, cement–sand mortar emerged as the superior choice, demonstrating the ability to sustain higher loads up to 25.60 kN, proving its cost-effectiveness and versatility for use in various locations within RC pipe joints. Conversely, non-shrinkage grout exhibited the lowest ultimate failure load, i.e., 21.50 kN, emphasizing the importance of material selection in enhancing the resilience and durability of urban infrastructure. A 3D finite element (FE) analysis was also employed to assess the effect of various factors on stress distribution and joint deformation. The findings revealed a 10% divergence between the experimental and numerical data regarding the ultimate load capacity of pipe joints, with experimental tests indicating a 25.60 kN ultimate load and numerical simulations showing a 23.27 kN ultimate load. Despite this discrepancy, the close concordance between the two sets of data underscores the utility of numerical simulations in predicting the behavior of pipe joints accurately. This study provides valuable insights into the selection and application of jointing materials in sewerage systems, aiming to improve the structural integrity and longevity of such critical infrastructure. Full article
Show Figures

Figure 1

Figure 1
<p>Steel reinforcement in conventional RC pipe.</p>
Full article ">Figure 2
<p>Pipe casting procedure; (<b>a</b>) steel fixing; (<b>b</b>) mold preparation with steel cage; (<b>c</b>) feeding concrete; (<b>d</b>) filled pipe in mold; (<b>e</b>) demolded concrete pipe; (<b>f</b>) pipe curing.</p>
Full article ">Figure 3
<p>Experimental setup for joint shear pipe testing.</p>
Full article ">Figure 4
<p>Schematic diagram of shear joint test.</p>
Full article ">Figure 5
<p>Numerical simulation of an RC pipe joint.</p>
Full article ">Figure 6
<p>Constitutive relationships (<b>a</b>) Hordijk model for tension of concrete; (<b>b</b>) Thorenfeldt model for compression of concrete; (<b>c</b>) Campbell model for compression of mortar; (<b>d</b>) Campbell model for shear of mortar.</p>
Full article ">Figure 7
<p>FEM pipes for joint shear test concrete modeling, with steel reinforcement meshing.</p>
Full article ">Figure 8
<p>Load-deflection curves of pipe joint tests.</p>
Full article ">Figure 9
<p>Schematic hairline crack location on pipe joint (<b>a</b>) upper side of the bell and (<b>b</b>) inclined side.</p>
Full article ">Figure 10
<p>Joint failure behavior of pipes with (<b>a</b>) non-shrinkage grout, and (<b>b</b>) cement–sand mortar.</p>
Full article ">Figure 11
<p>Validation of the numerical model with experimental results.</p>
Full article ">Figure 12
<p>Comparison of Load-Deflection Curves.</p>
Full article ">Figure 13
<p>Stress distribution on the pipe model.</p>
Full article ">
15 pages, 4028 KiB  
Article
Corrosion of Steel Rebars in Construction Materials with Reinforced Pervious Concrete
by Rosendo Lerma Villa, José Luis Reyes Araiza, José de Jesús Pérez Bueno, Alejandro Manzano-Ramírez and Maria Luisa Mendoza López
Infrastructures 2024, 9(4), 68; https://doi.org/10.3390/infrastructures9040068 - 1 Apr 2024
Viewed by 1751
Abstract
Pervious concrete has great potential for use in many practical applications as a part of urban facilities that can add value through water harvesting and mitigating severe damage from floods. The construction and agricultural industries can take direct advantage of pervious concrete’s characteristics [...] Read more.
Pervious concrete has great potential for use in many practical applications as a part of urban facilities that can add value through water harvesting and mitigating severe damage from floods. The construction and agricultural industries can take direct advantage of pervious concrete’s characteristics when water is a key factor included in projects as part of the useful life of a facility. Pervious concrete also has applications in vertical constructions, fountains, and pedestrian crossings. This work evidences that pervious concrete’s corrosion current increases with increasing aggregate size. Also, corrosion is a factor to consider only when steel pieces are immersed, aggravated by the presence of chlorine, but it drains water and does not retain moisture. Steel-reinforced pervious concrete was studied, and the grain size of the inert material and the corrosion process parameters were investigated. The electrochemical frequency modulation technique is proposed as a suitable test for a fast, reproducible assessment which, without damaging reinforced cement structures, particularly pervious concrete, indicates a trend of increasing corrosion current density as the size of the aggregate increases or density diminishes. Full article
Show Figures

Figure 1

Figure 1
<p>The materials used for manufacturing concrete of the types (<b>a</b>) conventional and (<b>b</b>) pervious.</p>
Full article ">Figure 2
<p>Geometry of the specimen.</p>
Full article ">Figure 3
<p>(<b>a</b>) Concrete mixer, (<b>b</b>) concrete cylinder specimens during setting, and (<b>c</b>) pervious and (<b>d</b>) conventional concrete specimens after curing for 28 days.</p>
Full article ">Figure 4
<p>Variable-load permeameter.</p>
Full article ">Figure 5
<p>(<b>a</b>) Aqueous solution of NaCl 5 wt%. (<b>b</b>) Specimens of conventional and pervious concrete immersed in a saline solution. (<b>c</b>) Conventional and (<b>d</b>) pervious concrete cylinder test specimens immersed in saline solution. The configuration of the (1) working, (2) counter, and (3) reference electrodes.</p>
Full article ">Figure 6
<p>Condition of specimens cut in halves with reinforcing steel rods after corrosion tests. (<b>a</b>) From left to right, No.4, 1/4″, 3/8″, 1/2″, 5/8″, and 3/4″. (<b>b</b>) Specimens No. 4 and Ref. 1 for comparison. (<b>c</b>) Zoom-in of void leaving the rebar naked that was frequent in pervious concrete specimens.</p>
Full article ">Figure 7
<p>Images of cylindrical test tubes before and after tests with conventional concrete (<b>a</b>,<b>b</b>), and pervious concrete (<b>c</b>,<b>d</b>), respectively.</p>
Full article ">Figure 8
<p>Specimens of permeable concrete were tested using the electrochemical frequency modulation technique. Two references using conventional concrete are shown.</p>
Full article ">Figure 9
<p>Specimens of permeable concrete were tested using the Tafel curve technique. Two references using conventional concrete are shown.</p>
Full article ">Figure 10
<p>Corrosion rates obtained using EFM and Tafel curves.</p>
Full article ">
16 pages, 5200 KiB  
Article
Experimental and Numerical Evaluation of Equivalent Stress Intensity Factor Models under Mixed-Mode (I+II) Loading
by Estefanía Gómez-Gamboa, Jorge Guillermo Díaz-Rodríguez, Jairo Andrés Mantilla-Villalobos, Oscar Rodolfo Bohórquez-Becerra and Manuel del Jesús Martínez
Infrastructures 2024, 9(3), 45; https://doi.org/10.3390/infrastructures9030045 - 1 Mar 2024
Cited by 4 | Viewed by 2346
Abstract
This study determines the equivalent stress intensity factor (SIF) model that best fits the experimental behavior of low-carbon steel under mixed modes (I and II). The study assessed Tanaka, Richard, and Pook’s equivalent SIF models. The theoretical values used for [...] Read more.
This study determines the equivalent stress intensity factor (SIF) model that best fits the experimental behavior of low-carbon steel under mixed modes (I and II). The study assessed Tanaka, Richard, and Pook’s equivalent SIF models. The theoretical values used for comparison correspond to the experimental results in a modified C(T) geometry by machining a hole ahead of the crack tip subjected to fatigue loads with a load ratio of R = 0.1. The comparison involved the SIF for six experimental points and the values computed through the numerical simulation. The Paris, Klesnil, and Modified Forman–Newman crack growth models were used with each equivalent SIF to analyze the prediction in the estimated number of cycles. The Klesnil model showed the closest prediction since the error between the calculated and experimentally recorded number of cycles is the lowest. However, the material behavior reflects a reduced crack propagation rate attributed to plasticity in the crack tip. The results suggest that Asaro equivalent SIF conservatively estimates the element lifespan with increasing errors from 2.3% at the start of growth to 27% at the end of the calculation. This study sheds light on the accuracy and limitations of different equivalent SIF models, providing valuable insights for structural integrity assessments in engineering applications. Full article
Show Figures

Figure 1

Figure 1
<p>Example of tubular specimen subjected to fully inversed axial (F) and torque (T) loads showing three opening modes. Adapted from [<a href="#B3-infrastructures-09-00045" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>Comparison of crack propagation models: Paris, Klesnil, and Modified Forman–Newman.</p>
Full article ">Figure 3
<p>Modified C(T) specimen; (<b>a</b>) dimensions in [mm] accordance with ASTM E647, (<b>b</b>) location of six reported intervals named 0 to e [<a href="#B30-infrastructures-09-00045" class="html-bibr">30</a>].</p>
Full article ">Figure 4
<p>Mesh details for the modified C(T) sample achieved by; (<b>a</b>) FEM; (<b>b</b>) BEM.</p>
Full article ">Figure 5
<p>Estimated cycles with the Paris rule; (<b>a</b>) Numerical vs. dN interval; (<b>b</b>) Mean percentage error concerning each interval of dN.</p>
Full article ">Figure 6
<p>Estimated cycles with the Modified Forman–Newman model; (<b>a</b>) Numerical vs. dN interval; (<b>b</b>) Mean percentage error with respect to each interval of dN.</p>
Full article ">Figure 7
<p>Estimated cycles with the Klesnil model; (<b>a</b>) Numerical vs. dN interval; (<b>b</b>) Mean percentage error concerning each interval of dN.</p>
Full article ">Figure 8
<p>Fatigue crack growth using K<sub>eq</sub> Asaro.</p>
Full article ">Figure 9
<p>Crack growth models using Asaro–Klesnil.</p>
Full article ">Figure 10
<p>(<b>a</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <msub> <mrow> <mi mathvariant="normal">K</mi> </mrow> <mrow> <mi mathvariant="normal">I</mi> </mrow> </msub> <mfenced open="[" close="]" separators="|"> <mrow> <mi>MPa</mi> <mo>⋅</mo> <msqrt> <mi mathvariant="normal">m</mi> </msqrt> </mrow> </mfenced> <mrow> <mo> </mo> <mi>vs</mi> </mrow> <mo>.</mo> </mrow> </semantics></math> a [mm]; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <msub> <mrow> <mi mathvariant="normal">K</mi> </mrow> <mrow> <mi>II</mi> </mrow> </msub> <mfenced open="[" close="]" separators="|"> <mrow> <mi>MPa</mi> <mo>⋅</mo> <msqrt> <mi mathvariant="normal">m</mi> </msqrt> </mrow> </mfenced> <mrow> <mo> </mo> <mi>vs</mi> </mrow> <mo>.</mo> </mrow> </semantics></math> a [mm]; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <msub> <mrow> <mi mathvariant="normal">K</mi> </mrow> <mrow> <mi>asaro</mi> </mrow> </msub> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mfenced open="[" close="]" separators="|"> <mrow> <mi>MPa</mi> <mo>⋅</mo> <msqrt> <mi mathvariant="normal">m</mi> </msqrt> </mrow> </mfenced> <mrow> <mo> </mo> <mi>vs</mi> </mrow> <mo>.</mo> </mrow> </semantics></math> a [mm]; (<b>d</b>) <math display="inline"><semantics> <mrow> <mrow> <mi mathvariant="normal">a</mi> <mo> </mo> </mrow> <mfenced open="[" close="]" separators="|"> <mrow> <mi mathvariant="normal">m</mi> </mrow> </mfenced> <mrow> <mo> </mo> <mi>vs</mi> </mrow> <mo>.</mo> <mrow> <mo> </mo> <mi mathvariant="normal">N</mi> <mo> </mo> </mrow> <mfenced open="[" close="]" separators="|"> <mrow> <mi>cycles</mi> </mrow> </mfenced> <mo>.</mo> <mo> </mo> <mo>(</mo> </mrow> </semantics></math><b>e</b>) <math display="inline"><semantics> <mrow> <mfrac> <mrow> <mi>da</mi> </mrow> <mrow> <mi>dN</mi> </mrow> </mfrac> <mfenced open="[" close="]" separators="|"> <mrow> <mfrac> <mrow> <mi>mm</mi> </mrow> <mrow> <mi>cycle</mi> </mrow> </mfrac> </mrow> </mfenced> <mi>vs</mi> <mo>.</mo> <mi mathvariant="sans-serif">Δ</mi> <msub> <mrow> <mi mathvariant="normal">K</mi> </mrow> <mrow> <mi>asaro</mi> </mrow> </msub> <mfenced open="[" close="]" separators="|"> <mrow> <mi>MPa</mi> <mo>⋅</mo> <msqrt> <mi mathvariant="normal">m</mi> </msqrt> </mrow> </mfenced> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Crack growth implementing mixed mode constants; (<b>a</b>) <math display="inline"><semantics> <mrow> <mfrac> <mrow> <mi>da</mi> </mrow> <mrow> <mi>dN</mi> </mrow> </mfrac> <mfenced open="[" close="]" separators="|"> <mrow> <mfrac> <mrow> <mi>mm</mi> </mrow> <mrow> <mi>cycle</mi> </mrow> </mfrac> </mrow> </mfenced> <mrow> <mi>vs</mi> <mo>.</mo> <mo> </mo> </mrow> <mi mathvariant="sans-serif">Δ</mi> <msub> <mrow> <mi mathvariant="normal">K</mi> </mrow> <mrow> <mi>asaro</mi> </mrow> </msub> <mfenced open="[" close="]" separators="|"> <mrow> <mi>MPa</mi> <mo>⋅</mo> <msqrt> <mi mathvariant="normal">m</mi> </msqrt> </mrow> </mfenced> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mrow> <mi mathvariant="normal">a</mi> <mo> </mo> </mrow> <mfenced open="[" close="]" separators="|"> <mrow> <mi mathvariant="normal">m</mi> </mrow> </mfenced> <mrow> <mi>vs</mi> <mo>.</mo> <mo> </mo> </mrow> <mrow> <mi mathvariant="normal">N</mi> <mo> </mo> </mrow> <mfenced open="[" close="]" separators="|"> <mrow> <mi>cycles</mi> </mrow> </mfenced> </mrow> </semantics></math>.</p>
Full article ">
13 pages, 7907 KiB  
Article
Fatigue Characteristics of Steel–Concrete Composite Beams
by Ayman El-Zohairy, Hani Salim, Hesham Shaaban and Mahmoud T. Nawar
Infrastructures 2024, 9(2), 29; https://doi.org/10.3390/infrastructures9020029 - 4 Feb 2024
Viewed by 1920
Abstract
Fatigue in steel–concrete composite beams can result from cyclic loading, causing stress fluctuations that may lead to cumulative damage and eventual failure over an extended period. In this paper, the experimental findings from fatigue loading tests on composite beams with various arrangements are [...] Read more.
Fatigue in steel–concrete composite beams can result from cyclic loading, causing stress fluctuations that may lead to cumulative damage and eventual failure over an extended period. In this paper, the experimental findings from fatigue loading tests on composite beams with various arrangements are presented. Fatigue tests were performed up to 1,000,000 cycles using four-point loading, encompassing various ranges of shear stress at a consistent amplitude. Additionally, the effects of external post-tensioning and the strength of the shear connection were investigated. Static tests were run until failure to assess the enduring strength of the specimens subjected to fatigue. The cyclic mid-span deflections, slippages, and strains were measured during the testing. Based on the experimental findings, it was found that the damage region that the shear studs caused in the concrete slab, which resulted in a reduction in stiffness within the shear connection, grew as the loading cycles increased, leading to an increase in residual deflections and plastic slippages. Controlling the longitudinal fatigue cracks in the concrete slab was largely dependent on the strength of the shear connection between the steel beams and concrete slabs. Moreover, the applied fatigue loading range affected the propagation and distribution of fatigue cracks in the concrete slab. The strains in different parts of the composite specimens were significantly reduced by applying the external post-tensioning. With no signs of distress at the anchors, the tendons displayed excellent fatigue performance. Full article
Show Figures

Figure 1

Figure 1
<p>Dimensions of the composite specimens (mm). (<b>a</b>) Specimens RSB 1, FSB 2 and FSB 3. (<b>b</b>) Specimen FSB 4. (<b>c</b>) Specimen FSB 5. (<b>d</b>) Specimens PRSB 6 and FPSB 7.</p>
Full article ">Figure 1 Cont.
<p>Dimensions of the composite specimens (mm). (<b>a</b>) Specimens RSB 1, FSB 2 and FSB 3. (<b>b</b>) Specimen FSB 4. (<b>c</b>) Specimen FSB 5. (<b>d</b>) Specimens PRSB 6 and FPSB 7.</p>
Full article ">Figure 2
<p>Post-tensioning system. (<b>a</b>) Schematic layout. (<b>b</b>) On-site photo.</p>
Full article ">Figure 3
<p>Test setup (mm). (<b>a</b>) Fatigue test. (<b>b</b>) Static test.</p>
Full article ">Figure 4
<p>Instrumentations. (<b>a</b>) Strain gauges and LVDTs layout. (<b>b</b>) Slippage LVDT and post-tensioning load cells. (<b>c</b>) Shear stud’s strain gauges.</p>
Full article ">Figure 5
<p>Development and dispersion of fatigue cracks in the concrete slabs (cycle counts are displayed in thousands). (<b>a</b>) Specimen FSB 2. (<b>b</b>) Specimen FSB 3. (<b>c</b>) Specimen FSB 4. (<b>d</b>) Specimen FSB 5. (<b>e</b>) Specimen PFSB 7.</p>
Full article ">Figure 5 Cont.
<p>Development and dispersion of fatigue cracks in the concrete slabs (cycle counts are displayed in thousands). (<b>a</b>) Specimen FSB 2. (<b>b</b>) Specimen FSB 3. (<b>c</b>) Specimen FSB 4. (<b>d</b>) Specimen FSB 5. (<b>e</b>) Specimen PFSB 7.</p>
Full article ">Figure 6
<p>Cyclic deformations. (<b>a</b>) Cyclic deflection. (<b>b</b>) Cyclic slippage.</p>
Full article ">Figure 7
<p>Cyclic strains in the shear connectors.</p>
Full article ">Figure 8
<p>Cyclic strains in the concrete slabs and steel beams. (<b>a</b>) Concrete slab. (<b>b</b>) Steel beam.</p>
Full article ">Figure 9
<p>Fatigue variations in the post-tensioning force.</p>
Full article ">Figure 10
<p>Static residual deformations. (<b>a</b>) Residual deflection. (<b>b</b>) Residual slippage.</p>
Full article ">Figure 11
<p>Static residual strains in the extreme fibers of the composite sections. (<b>a</b>) Concrete slab. (<b>b</b>) Steel beam.</p>
Full article ">Figure 12
<p>Static residual post-tensioning force.</p>
Full article ">
24 pages, 21460 KiB  
Article
Strength and Deformation of Concrete-Encased Grouting-Filled Steel Tubes Columns Exposed to Monotonic Quasi-Static Loading Conditions
by Ahlam A. Abbood, Nazar Oukaili, Abbas A. Allawi and George Wardeh
Infrastructures 2024, 9(2), 26; https://doi.org/10.3390/infrastructures9020026 - 1 Feb 2024
Cited by 1 | Viewed by 1851
Abstract
This study aimed to evaluate the effectiveness of a novel concrete-encased column (CE) using small circular steel tubes filled with cementitious grouting material (GFST) as the primary reinforcement instead of traditional steel bars. The research involved three different types of reinforcement: conventional steel [...] Read more.
This study aimed to evaluate the effectiveness of a novel concrete-encased column (CE) using small circular steel tubes filled with cementitious grouting material (GFST) as the primary reinforcement instead of traditional steel bars. The research involved three different types of reinforcement: conventional steel bars, concrete-filled steel tubes with 30% of the reinforcement ratio of steel bars, and concrete-filled steel tubes with the same reinforcement ratio as steel bars. Twenty-four circular concrete columns were tested and categorized into six groups based on the type of reinforcement employed. Each group comprised four columns, with one subjected to concentric axial load, two subjected to eccentric axial load (with eccentricities of 25 mm and 50 mm, respectively), and one tested under lateral flexural loads. To validate the experimental results, finite element (FE) analysis was conducted using ABAQUS software version 6.14. The experimental findings for concentric load reveal that columns with the second type of reinforcement, concrete-filled steel tubes with 30% of the reinforcement ratio of steel bars exhibited a failure load 19% lower than those with steel bars, while columns with the third type of reinforcement, concrete-filled steel tubes with the same reinforcement ratio as steel bars achieved a failure load 17% greater than the traditional steel bars. The FE analysis demonstrates good agreement with the experimental outcomes in terms of ultimate strength, deformation, and failure modes. Full article
Show Figures

Figure 1

Figure 1
<p>Cross-sections of CE-CFST columns [<a href="#B8-infrastructures-09-00026" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>Schematic of cross-section dimensions and reinforcement details of tested specimens.</p>
Full article ">Figure 3
<p>Steel loading heads and saddles were used in the test.</p>
Full article ">Figure 4
<p>Typical test setup and instrumentation for (<b>a</b>) column specimen; (<b>b</b>) beam specimen.</p>
Full article ">Figure 4 Cont.
<p>Typical test setup and instrumentation for (<b>a</b>) column specimen; (<b>b</b>) beam specimen.</p>
Full article ">Figure 5
<p>The mesh size of all parts of the tested specimen.</p>
Full article ">Figure 6
<p>FE boundary conditions and applied load were used in the analysis.</p>
Full article ">Figure 7
<p>Stress–strain curves of material (<b>a</b>) unconfined concrete in compression; (<b>b</b>) confined grouting material in compression; (<b>c</b>) concrete and grouting material in tension; (<b>d</b>) stress–strain for steel tube.</p>
Full article ">Figure 8
<p>Ultimate load of all tested specimens under various loading conditions.</p>
Full article ">Figure 9
<p>Effect of spiral spacing on ultimate load for specimens tested under various loading conditions.</p>
Full article ">Figure 10
<p>Experimental and numerical load–axial deformation curves of the concentric specimens.</p>
Full article ">Figure 11
<p>Experimental and numerical load–lateral deformation curves of eccentric specimens.</p>
Full article ">Figure 12
<p>Experimental and numerical load–deflection curves of flexural specimens.</p>
Full article ">Figure 13
<p>Ductility index of tested specimens under eccentric and flexural loads.</p>
Full article ">Figure 14
<p>Experimental and numerical failure mode of specimens under concentric load.</p>
Full article ">Figure 14 Cont.
<p>Experimental and numerical failure mode of specimens under concentric load.</p>
Full article ">Figure 15
<p>Experimental and numerical failure mode of specimens under 25 mm eccentric load.</p>
Full article ">Figure 15 Cont.
<p>Experimental and numerical failure mode of specimens under 25 mm eccentric load.</p>
Full article ">Figure 16
<p>Experimental and numerical failure mode of specimens under 50 mm eccentric load.</p>
Full article ">Figure 16 Cont.
<p>Experimental and numerical failure mode of specimens under 50 mm eccentric load.</p>
Full article ">Figure 17
<p>Experimental and numerical failure mode of specimens under flexural load.</p>
Full article ">Figure 17 Cont.
<p>Experimental and numerical failure mode of specimens under flexural load.</p>
Full article ">Figure 18
<p>Experimental and numerical strength interaction diagrams for the six groups.</p>
Full article ">Figure 18 Cont.
<p>Experimental and numerical strength interaction diagrams for the six groups.</p>
Full article ">

Other

Jump to: Editorial, Research

1 pages, 127 KiB  
Correction
Correction: Pour et al. Enhancing Flexural Strength of RC Beams with Different Steel–Glass Fiber-Reinforced Polymer Composite Laminate Configurations: Experimental and Analytical Approach. Infrastructures 2024, 9, 73
by Arash K. Pour, Mehrdad Karami and Moses Karakouzian
Infrastructures 2024, 9(7), 111; https://doi.org/10.3390/infrastructures9070111 - 15 Jul 2024
Viewed by 489
Abstract
In the published publication [...] Full article
14 pages, 5911 KiB  
Technical Note
Practical Aspects of Correlation Analysis of Compressive Strength from Destructive and Non-Destructive Methods in Different Directions
by Baitollah Badarloo and Petr Lehner
Infrastructures 2023, 8(11), 155; https://doi.org/10.3390/infrastructures8110155 - 24 Oct 2023
Viewed by 2133
Abstract
The research presented here demonstrates the practical aspects of the numerical correlation of the results of the compressive strength test. The destructive test (DT) in a hydraulic press and the non-destructive test (NDT) using a Schmidt hammer in several process variations were evaluated. [...] Read more.
The research presented here demonstrates the practical aspects of the numerical correlation of the results of the compressive strength test. The destructive test (DT) in a hydraulic press and the non-destructive test (NDT) using a Schmidt hammer in several process variations were evaluated. The aim was to evaluate the real differences between the tool supplier’s curve and testing. Therefore, 150 concrete cube specimens with an edge length of 150 mm were produced using a mixture of three types of concrete classes: C30, C35, and C40. The test was carried out 7 and 28 days of age of the concrete. The Schmidt hammer test was carried out in horizontal (θ = 0) and vertical (θ = 90) directions and using a series of 10 measurements. Furthermore, the tests were performed in two sets: first, the sample was placed on the ground, and second, under a hydraulic jack with a load of 50% of the maximum bearing capacity of specific concrete. Then, regression analysis was performed on the data sets to establish linear mathematical relationships between compressive strength and number of bounces. The results showed that the correlation between the DT and NDT tests has a high value for each group, but the correlation equations are different and must be taken into account. Full article
Show Figures

Figure 1

Figure 1
<p>Steps to use a Schmidt hammer: (<b>a</b>) placing the calibration element, (<b>b</b>) testing vertically, (<b>c</b>) testing horizontally.</p>
Full article ">Figure 2
<p>Correlation of the rebound index of the horizontal test and compressive strength at 7 days for concrete C30.</p>
Full article ">Figure 3
<p>Correlation of the rebound index from the horizontal test and compressive strength at 7 days for concrete C35.</p>
Full article ">Figure 4
<p>Correlation of the rebound index of the horizontal test and compressive strength at 7 days for concrete C40.</p>
Full article ">Figure 5
<p>Correlation of the rebound index of the vertical test and compressive strength at 7 days for concrete C30.</p>
Full article ">Figure 6
<p>Correlation of the rebound index of the vertical test and compressive strength at 7 days for concrete C35.</p>
Full article ">Figure 7
<p>Correlation of the rebound index of the vertical test and compressive strength at 7 days for concrete C40.</p>
Full article ">Figure 8
<p>Correlation of the rebound index from the horizontal test and compressive strength at 28 days for concrete C30.</p>
Full article ">Figure 9
<p>Correlation of the rebound index of the horizontal test and compressive strength at 28 days for concrete C35.</p>
Full article ">Figure 10
<p>Correlation of the rebound index from the horizontal test and compressive strength at 28 days for concrete C40.</p>
Full article ">Figure 11
<p>Correlation of the rebound index of the vertical test and compressive strength at 28 days for concrete C30.</p>
Full article ">Figure 12
<p>Correlation of the rebound index of the vertical test and compressive strength at 28 days for concrete C35.</p>
Full article ">Figure 13
<p>Correlation of the rebound index of the vertical test and compressive strength at 28 days for concrete C40.</p>
Full article ">
Back to TopTop