Next Article in Journal
Dynamic Modal Analyses of Building Structures Employing Site-Specific Response Spectra Versus Code Response Spectrum Models
Next Article in Special Issue
Development of Prediction Model for Rutting Depth Using Artificial Neural Network
Previous Article in Journal
Review of Accelerated Bridge Construction Systems for Bridge Superstructures and Their Adaptability for Cold Weather
Previous Article in Special Issue
A Three-Phase Model to Evaluate Effects of Phase Diffusivity and Volume Fraction upon the Crack Propagation in Concrete Subjected to External Sulphate Attack
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessing Effectiveness of Shape Memory Alloys on the Response of Bolted T-Stub Connections Subjected to Cyclic Loading

1
Department of Civil Engineering and Water Engineering, University of Laval, Quebec, QC G1V 0A6, Canada
2
Department of Architectural Engineering, Pennsylvania State University, State College, PA 16801, USA
3
Department of Civil, Environmental and Infrastructure Engineering, George Mason University, Fairfax, VA 22030, USA
4
Department of Marine Engineering, Amirkabir University of Technology (Tehran Polytechnic), Tehran 1591634311, Iran
*
Author to whom correspondence should be addressed.
CivilEng 2023, 4(1), 105-133; https://doi.org/10.3390/civileng4010008
Submission received: 29 September 2022 / Revised: 21 January 2023 / Accepted: 24 January 2023 / Published: 30 January 2023
(This article belongs to the Special Issue Next Generation Infrastructure)

Abstract

:
This study presents finite element analysis of double split tee (DST) connections with high-strength steel bolts and coupled split tee sections, to evaluate various cyclic response parameters and elements. The investigation included quantifying connection behavior and hysteretic response, damage indexes, and failure modes. Over 40 specimens were simulated in ABAQUS under cyclic loading, including shape memory alloy (SMA)-built specimens. In the post-analysis phase, the T-stub thickness, the T-stub yield strength, the bolt preload and bolt number, and the stiffener type and stiffener material for the most significant parts of the DST connection were calculated. Simulation results showed that a lower ultimate moment yielded fewer needed stem bolts. The energy dissipation (ED) capacity increased as the horizontal distance between the stem bolts decreased. Additionally, increasing the strength of the bolt and T-stub by 15% resulted in a 3.86% increase in residual displacement (RD) for the bolt and a 1.73% decrease in residual displacement for the T-stub. T-stub stiffeners enhanced ED capacity by 31.7%. SMA materials were vulnerable to mode 1 failure when used in T-stubs, bolts, or stiffeners. However, the use of SMA increased the rate of energy dissipation. Adding stiffeners to the T-stubs altered the failure indexes and improved the pattern of failure modes. In addition, stiffeners decreased the rupture and pressure indexes. As a result, the failure index of a T-stub shifted from brittle failure to ductile failure.

1. Introduction

In traditional seismic design, connections between strong columns and weak beams have been used to achieve the desired modifications to stability and cyclic behavior [1,2,3,4]. The connections also protect against soft-story occurrence via the energy dissipation (ED) mechanism [5]. However, detailing reinforcement for structural joints that must meet full-strength requirements with an over-strength factor is reported to be expensive [6,7,8,9].
Conventional T-stub connections involve welding or bolting T-stubs to the primary structural members [10]. The connection provides high flexural resistance. Both tensile and shear behaviors can be studied by deforming the top T-stub axially and transversely [11,12]. Figure 1 is an overarching illustration of how a load affects a bolted steel connection and a T-stub.
As shown in Figure 1, it is possible to develop a prying action between the T-stub flange and the support base. By preloading the T-stub bolts, a contact stress is created between the T-stub flange and the support base. Atasoy [13] and Kulak et al. [14] note that when external loads are applied to the T-stub web, a contact pressure might be generated between the flange edge and the base. Figure 1 also shows how such contact stress would overload the T-stub bolts with a prying action force, Q, based on the flexural rigidity of the T-stub base flange. As a result of this force, the T-stub connector bolts are at risk of premature failure [15].
There is also an uncertainty about the position of the prying action force in the current state of the art regarding T-stub connectors. The force is assumed to act on the side line of the T-stub in a wide range of analytical models, such as those in [15,16,17,18,19,20,21]. AISC [16,22] and ASCE [23] have adopted this assumption.
The strength of tension zones in column flanges, end-plates, and bolts has received more attention [17]. The American Institute of Steel Construction (AISC) [24] and the Eurocode 3 [25] Standards provide analysis methods that serve as starting points from which seismic design enhancements to end-plate and T-stub connections could be derived.
Several analytical studies [18,26,27,28,29,30,31] verified static equilibrium and the application of the beam theory to the T-stub flange, and several experiments [32,33,34] were conducted to establish the force–displacement behavior. The T-stub connections’ rotational capacity was identified and evaluated using post-limit behavior [9,32,33,34,35], revealing the connections’ resistance and stiffness.
Abidelah et al. [36] analyzed the impact of bolts on T-stub performance and looked into the causes of failure. The functionality of a welded T-stub was studied by Barata et al. [37]. They discovered that when the temperature increased, the T-fire stub’s resistance and initial stiffness decreased, and its failure modes shifted.
Additionally, a three-dimensional FE model was used to study the structural behavior of the moment resisting connections and the formation of the plastic hinges in steel beam–column connections under earthquake load, and the stiffness and resistance predictions were presented in detail, with a model validated via experimental results [38,39].
For bolted T-stub connections including welded plates, the welding method is very crucial for ensuring ductile behavior of the connection [40]. T-stubs are sensitive to the minor axis of the column in out-of-plane bending [41].
High ductility in a moment resisting frame (MRF) was achieved by fabricating specimens with both unstiffened and stiffened end plate connections, with the latter using bolt extensions [42]. Diagrams illustrating the bending moment and deflection of an MRF under seismic load are shown in Figure 2.
Ozkilic [44,45,46,47] evaluated behavior of stiffened extended end-plate connections with large bolts, application of extended end-plate connections for replaceable shear links, effects of yield line mechanisms of four bolted extended end-plate connections on frame response, and impacts of bolt diameter and end-plate thickness on capacity of connection and failure modes under cyclic and monotonic loading.
Guidelines and standards have been established by Mistakidis et al. [48], Gebbeken et al. [49], Sherbourne and Bahaari [50], Bursi and Jaspart [32,51,52,53,54], Hantouche et al. [55,56,57,58,59], Lemonis and Gantes [60,61,62,63], and Swanson et al. [33]. Elementary T-stubs’ behavior was investigated by Soltani et al. [64] using finite element analysis. All possible modes of failure were accounted for by considering a wide range of geometry and strength characteristics. There were examples of both identical and dissimilar T-stub configurations among the specimens. To determine the full force–displacement curve of bolted connections, Fernandez-Ceniceros [65,66,67] has conducted extensive studies. Precise estimation of link behavior was achieved using the hybridization of FE modeling with soft-computing approaches [68].
Bravo and Herrera [69], Girao Coelho et al. [42], Hantouche et al. [59], and Herrera et al. [70] have investigated the behavior of bolted steel connections in an effort to determine their failure modes and fracture zones. As a result, bridge piers are ideal locations for conducting scour experiments [71]. The nonlinear behavior of a T-stub join under impact stresses was also established in a study by Ribeiro et al. [72]. The force–displacement results were improved by using a ductile failure criterion, and it was found that the strain rate had an effect on the material’s behavior. Different dynamic loads had little effect on the force–displacement response, despite the fact that they all applied a sizable impetus.
The finding was that an increase in bolt diameter results in an increase in plastic capacity as well as an increase in gage distance. Additionally, if the thickness of the weld increases, the plastic capacity will increase as well. Moreover, strain hardening affected the capacity of T-stub connections [73,74,75,76].
The bolted angle connection’s moment rotation capacity and influencing parameter were analyzed by Saedi Daryan et al. When the shear hardness increased, the lateral drift decreased, and the frame stability increased [77].
Numerical models were developed to study the performance of the post-tensioned (PT) self-centering steel connection in the context of an intelligent seismic system. Different types of bolted angles with associated stiffeners (Shiravand et al. [78] and Abdollahzadeh et al. [79]) and SMA material [80,81] were examined for their impact on structural member behavior.
In keeping with this discovery, Asada et al. [82] described a method for increasing the plasticity of welded wide flange beam to column connections by using high-yield bolts and welding complementary H-section haunches.
Sebbagh et al. [83] found that if the flange is monotonically loaded, adding an outer row of bolts is unlikely to have any significant effect until it begins to yield. When cyclic loads are applied, the added row of bolts prevents stiffness from decreasing and can provide more than 80% more strength under cyclic loads. As a result of cyclic loading, the outer bolts take on practically all the load applied during the unloading phase while simultaneously participating in the load resistance during loading. With the addition of the outer bolts, the end part of the T-stub plate under cyclic loads is prevented from uplift, and prying forces are avoided, while the inner bolts undergo about 40% less elongation during loading and unloading.
The stainless steel T-stub to the square hollow section bolted connection was numerically studied [84]. It was shown that by increasing the flange thickness from 2 mm to 7 mm, the connection capacity was increased by 88%. Bolts positioned far apart from the stem of the T stub exhibited a ripping or prying action. Additionally, thicker flanges were less susceptible to prying. With the diameter of the bolt increasing from 12 mm to 16 mm, the capacity increased by 21.4%.
Designing the flange bearing resistance against the membrane shear component can increase the T-stub’s ductility at extremely large deformation. Finite element calculations estimate the hole’s design force as 0.2 times the bolt’s tensile resistance. Constructional flaws, such as web misalignment and flange bowing, only affect bolt action. These discrepancies are minimal and can be ignored for the examined flaw [85].
Five beam-to-column joints, made of austenitic stainless steel, were put through cyclic loading tests by Wang et al. [86]. The joints were connected using welded flange and bolted web connections. Cyclic tests were performed on three beam-to-column joints made of duplex stainless steel by Liu and Shi [87] using welded unreinforced flange-welded web connections, while experimental results from five austenitic stainless steel bolted extended end-plate beam-to-column joints were reported by Bu et al. [88]. Furthermore, Yuan et al. [89] performed a battery of cyclic experiments on stainless steel T-stubs and developed a new hysteretic model. The hysteretic performance of stainless steel end-plate beam-to-column joints has not been thoroughly studied, and only five joints of austenitic grade S31608 were tested under cyclic pressure.
Zhao et al. [90] found that grades S385 and S440 specimens were more malleable than S690 specimens causing brittle fracture. Kong and Kim [10] provided the results of finite element simulations, with a particular emphasis on the anticipation of the first-stage hardness and final-stage moment for T-stub joints. To enhance the moment and initial stiffness capability, a novel model was presented, and the connection collapse mechanisms were investigated.
Sun et al. [91] demonstrated the dampening effect of the heat affected zone (HAZ) by applying shear and axial loads to a T-stub joint. As a result, as the dispersion radius of HAZ was extended, and the initial yield resistance of the T-stub represented a linear distance [91].
Using MRFs, Wang et al. [92] studied the operation of the blind-bolt Hollobolt. Initial stiffness, malleability, and strength of the bolted steel connection were all measured in the analysis. Further, by contrasting the flat angle and the shank diameter of the bolt, it was discovered that the flange thickness had a direct bearing on the stiffness or strength of the blind-bolted connections.
The study in this paper presents the structural responses of a DST connection when loaded laterally. Thus, a few parameters were chosen to examine the connection’s response modification. When the connection is subjected to seismic loads, a number of factors must be modified, including the behavioral requirements of the bolt and T-stub, the material of the effective section (from steel to SMA), and the geometry of the stiffener. The fundamental goal of this research was to establish the correct DST connection behavior. Accordingly, the proposed damage indexes, failure modes, and behavioral curves were evaluated so that the structural response of a bolted steel connection may be predicted, allowing for a more robust and efficient overall structural system to be designed. The significance of the study is to assess the performance of DST connections that include stiffeners and shape memory alloys (SMA) because supplementary elements are increasingly employed in steel connections to improve and alter their behavior. The paper concludes by outlining the necessary steps and projected outcomes for improving the DST link.

2. Finite Element Models

The reference FE model consisted of a full-scale T-stub connection, calibrated by the experimental model (Yang et al. [93]). All other comparative models were developed by altering different parameters from the reference model. This connection is composed of the beam, column, double plates, continuity plate, bolts, and T-stub. The height of the column and length of beams is equal to 1 m. The beam has lateral support to avoid unwanted movement, overturning, and out-of-plane connection rotation during cyclic loading. The load is applied to the free end of the beam. In this research, the methodology employed included the stress amplitude and cycle (SAC) loading protocol [94]. This protocol consists of applying a specific level of stress and a specific number of cycles to a material and then evaluating its response. As for the loading phase, cyclic loading was applied to the tip of the beam section, and a static general analysis method was used.
The sections of the beam and column were H406 × 403 × 16 × 24 mm. The T-stub area was defined by beam flanges, column flanges, and prestressed bolts. The distance between bolts and hole arrangements was defined in the experimental models.
The model assumed that the sliding contact between the bolt shank and the bolt hole, the bolt head and the flange, and the bolt head and the rigid plate was finite, allowing for the simulation of massive slip. Normal behavior was assumed to be hard contact, while tangential behavior was simulated using the penalty friction method. In the literature for steel, stainless steel, and aluminum connections [38,92,95,96,97,98], the chosen co-efficient of friction was 0.3, which is within the range of 0.2–0.33.
The specification of the model parts is defined in Table 1, in which values of Fu and Fy for bolts, T-stub, beam, and column are presented. Fy and Fu are yield stress and ultimate stress, respectively.
The axisymmetric condition was established for modeling for computational efficiency. A decrease in the quantity of bolts and interactions was modelled for half of the beam-to-column connection. The length of the T-stub section, the stub column, and the midpoint of the beam were all constrained in the X and Z directions to ensure connection symmetry.
Figure 3 demonstrates a meshing overview of the solid elements. The eight-node 3D solid elements and the reduced integral (C3D8R) were used to deliver high precision at minimal computational cost [33,38,92,99]. Several mesh densities were tested, and eventually, a structured mesh was used with variable densities for each region of the model. In order to accurately capture their out-of-plane flexure and prevent the effect of shear locking, at least three elements were given through the thickness of the T-stub. According to [95,96,97,99], a flange’s thickness should have at least two elements across it for bending to be effective. To better simulate the interface between the bolt and the T-stub and to capture contact, a finer mesh was employed around the smaller region of the bolt [99].
Mesh sensitivity analysis was conducted to obtain an optimized mesh size for all elements of the T-stub connection. Several mesh sizes were studied to determine the size that provided the best accuracy. In the panel zone, which was subject to more stress and strain variations, and also for the T-stub sections and column bolts, which underwent significant yielding, finer meshes were specified.
The total numbers of nodes and elements were 32,760 and 19,572 with linear hexahedral elements (C3D8R), respectively. The column section’s mesh sizes at the surroundings of the bolt hole and bolt section were 1.8 and 3.6 cm, respectively. The beam’s mesh size adjacent to an area of the column was about 1.5 cm. In addition, plates and T-stubs were supposed to be 2.1 cm and 1.8 cm, respectively. Mesh sensitivity analysis was conducted to compare the stress value of the T-stub section or area as it provided an optimized meshing size.
The analysis had two steps; in the first step, the bolts were prestressed using the bolt load method (AISC 360-16(2016) and RCSC (2014)) [100,101]. In the second step, the cyclic loading was applied to the beam’s free end. The numerical modeling results were compared with six specimens experimentally [93], as shown in Table 2. The effective parameters were M0, Mmax, and θ max defining the reference moment of the T-stub, maximum moment, and maximum rotation, respectively. As a result, the maximum error of numerical modeling was approximately 3.47% compared with the experimental results. Figure 4 illustrates the force–displacement curve and Von Mises distribution for the experimental specimen (G310-T19-B400). The stresses concentrated on the T-stub through the beam rotation, and the critical stress and local deformation were revealed in beam flanges and bolts, respectively. Nu and Ex are the abbreviation forms of the numerical and experimental values, respectively.
The G310-T19-B400 model was considered the verification model (SP02) for further investigation. The parametric specifications for the developed models are presented in Table 3, where tf, Fyt, and Fyb are the T-stub’s thickness, the T-stub’s yield stress, and the bolt’s yield stress, respectively. L1 and L2 are the arrangement of bolt holes on the T-stub web, as depicted in Figure 5. All dimensions are in millimeters.
The SP01 to SP03 models showed a decrease in L2 value of ~50%. The SP04 to SP07 models were classified based on changing the number and material of bolts. For example, in the SP04 model, the bolts that connected the T-stub section to beam flanges were reduced from six to two in number. Based on the SP08 to SP13 models, a change in the T-stub’s yield stress and thickness or the bolt pretension load could be seen. The steel and SMA were equipped with either a bolt column or a T-stub, as shown in the SP14 and SP15 models. In the SP16 to SP19 models, the T-stub section was equipped with a triangular stiffener as the thickness and number differed.

The Capability of SMA Material

Shape memory alloy (SMA) has been used in seismic applications because it can withstand large loads and recover its original shape by unloading and heating. Utilizing SMA is related to self-centering and energy dissipation capacities in behavior (Wang et al. [103] and Wang and Zhu [104]). According to Pereiro-Barceló et al. [105], the reversible transitions between martensite and austenite identify this distinct property. The SMA likely recovers up to 8% strain in the mode of NiTi alloy because the austenite phase exhibits superplastic behavior (Chowdhury et al. [106]). NiTi material is imported into the ABAQUS based on the mechanical characteristics described in DesRoches et al. [102].

3. Analyzing the Moment–Rotation Curve

The parameters of the moment–rotation curve for 19 models are presented for understanding the behavior of bolted T-stub connection. These parameters include moment capacity, ED capacity, and RD, as shown in Table 4.
A downward trend in the Mmax, ED, and RD was revealed when there was an increasing trend in the value of L1. For instance, in the SP03 model, ED capacity was reduced compared with the SP02 model (reference model) by about 20%. Similar reductions for the SP06 and SP07 models were observed by a decrease in yield stress of the bolt. A leap in Fyt, tf, or bolt pretension led to enhancing the ED, Mmax, and RD values. For example, the values of ED, Mmax, and RD had increments of 1.5%, 3%, and 6%, respectively, where tf was increased to 23 mm.
A sharp drop in the Mmax and ED values in the SP15 model could be seen when the T-stub section was equipped with SMA. While an increment trend in the values could be seen when the T-stub was equipped with triangular stiffeners, the maximum ED increased by approximately 55%.

Assessment of Response Curve Variables

The response curve variables are provided by using the force–displacement curve as shown in Figure 6. Table 5 presents the response variable factors. The initial stiffness (Ki) of a bolted steel connection at a drift of 0.3% is defined as the secant stiffness and is calculated from the cyclic response. The residual stiffness or post-yield stiffness (Kr) of a bolted steel connection after the decompression has occurred is also shown, as well as the maximum force (Fmax) in the connection under lateral loading.
The values of Ki, Kr, and Fmax showed a downward trend by the change in Fyt and tf and also when the T-stub or bolt was equipped with SMA material.
The maximum value of Ki limited to the SP17 model was substantially increased by about 102% compared with the SP02 model, while the minimum value showed a 50% decrease for the SP14 model. The value of Kr fluctuated significantly. For example, for the values of the SP12 and SP04 models, the trend decreased by 51% and increased by 75%, respectively. In the SP17 model, Fmax showed an upward trend of 24%, although its lowest value had a decreasing trend of 11.5% compared with the SP02 model, where the number of beam bolts changed from six to two.
The use of SMA material for the T-stub section resulted in a reduction in the initial stiffness (Ki), residual stiffness or post-yield stiffness (Kr), and maximum force (Fmax) because of the inherent elastic, resilient, and reversible properties of the SMA material. However, it improved the flexibility and ductility of the DST connection.
As for stiffener-equipped models, these components affected the total response of the connections significantly. The Ki and Fmax values increased. Stiffeners enhanced the stiffness of the DST connection, and the connection would sustain 5% drift with negligible strength degradation.
Figure 7 depicts the structural behavior curves and Von Mises distributions for models SP14 and SP15. Consequently, in the SP15 model, a reduction in the strength could be noticeably seen compared with the SP14 model. Based on the Von-Mises distribution at a 4% drift, the SP14 model shows a significant focus on stress in comparison to the SP15 model. However, in the SP15 model, there is a sharp increase in the strain energy value that occurs when the drift exceeds 2%.
The ductility capacity and residual displacement of the SP14 model were more than 20 and 40%, respectively, compared with the SP15 model. This conclusion was derived from Figure 7b. These changes illustrated more effectiveness of the SMA column bolts than the SMA T-stub in enhancing the ductility and residual displacement of the bolted steel connection.
Figure 8 shows the stress–strain diagram of the SMA T-stub. In this case, the T-stub strain did not exceed 6%. Thus, the conclusion was that the behavior of the SMA materials was simulated correctly.

4. Failure Indexes

Since fractures or cracks were not explicitly modeled, response indexes could be used to determine the potential level of fracture. There are several commonly used indexes to assess potential fracture, including the pressure index, the Mises index, the triaxiality index (TI), the equivalent plastic strain index (PEEQ), and the rupture index (RI) [107]. An accumulation of high RI values indicated the location of the most potential fracture [108]. Initiating fractures could not be determined by RI. Bozkurt et al. [109,110,111] recommended using it to compare different configurations to determine the most likely to fracture. In many studies, researchers use the rupture index to evaluate the degree of damage to be experienced at beam-to-column connections [107,112,113,114,115].
The T-stub section had a significant role in the DST connection’s cyclic behavior to dissipate energy and affect the connection’s stiffness and strength.
Equations (1)–(3) show an example of damage index formulations of the T-stub from El-Tawil et al. [116].
PEEQ = 2 3   ε ij p ε ij p
where the PEEQ index is the ratio of the equivalent plastic strain by the yield strain and ε i j p is the plastic strain.
TI = PI MI
According to Equation (2), the triaxiality index ( TI ) is the ratio of hydrostatic stress ( σ m ) to the Mises stress ( σ ¯ ) . The pressure index ( PI ) is the hydrostatic stress ( σ m ) divided by the yield stress ( σ y ), and the Mises index ( MI ) is defined as the ratio of the Mises and yield stress. In Equation (3), the rupture index (RI) is the PEEQ divided by e ( 1.5 σ m σ ¯ ) .
RI = PEEQ exp ( 1.5 σ m σ ¯ )
Additional indexes such as the pressure index, the strain index, and a mix of them were calculated to compare and assess the behavior of the connection system, as shown in Table 6. The stress and strain values were normalized according to the yield stress and strain of the T-stub section.
The pressure index indicates a failure possibility in the area of the T-stub. The existence of hydrostatic stress in this area increased the likelihood of sudden failure and fracture of the T-stub. Decreasing the number of connected bolts to the beam reduced the T-stub fillet’s pressure index and sudden failure possibility. Moreover, a decrease in the spaces between bolts increased the pressure index. It caused improvement in the distribution of ultimate stress and strain on the T-stub section, and finally, the failure of the T-stub section became flexible.
The equivalent plastic strain index (PEEQ) indicates the flexibility of a specific section of the T-stub. The rupture index is the ratio of triaxiality stress (triple ratio). The medium range of triaxiality stress was between values of 0.75 and 1.5 [116]. When the fractural behavior of models became brittle, the triaxiality stress became more than 1.5 or lower than 0.75.
This research compared and evaluated the flange and web area of the T-stub or T-stub fillet index. Increasing the yield stress in the SP08, SP10, and SP12 models changed the value of the rupture index from 1.5 to 1.41, 1.53, and 1.55, respectively. In other words, the fractural behavior of the T-stub became brittle. However, according to the pressure index, increasing the thickness of T-stub and preload of bolts indicated better unexpected fractural behavior in mentioned models.
The rupture index, pressure index, and triaxiality stress index were reduced significantly in models equipped with stiffeners. The reduction of the rupture index of the SP17 model was about 1.05, according to the reference index value (1.5). As a result, the stiffeners improved the sudden fractural behavior of the T-stub section, and the failure of the T-stub connection became flexible.

Failure Modes

This section covers the Eurocode 3 [25] approach. A simple bolted T-stub can fail in accordance with three possible collapse cases, in which the occurrence of a given case typology is governed by the parameter β Rd as shown in Equation (4):
β Rd = 4 M f . Rd 2 B Rd m
In the above equation, Mfrd is the design flexural resistance of the T-stub flanges, BRd is the design axial resistance of the single bolt, m is the distance between the bolt axis and the section corresponding to the flange-to-web fillet where the formation of a plastic hinge can be expected, and n is the distance between the bolt axis and the location of the prying force.
In the first case, four plastic hinges were used. Two hinges at the bolt axes assisted with two hinges at the flange to web connection. The value of the design resistance is defined as in Equation (5):
F 1 . Rd = 4 M f . Rd m
The second case was associated with the formation of two plastic hinges at the sections dealing with the flange-to-web connection and by the fracture of the bolts. The design resistance is given by Equation (6):
F 2 . Rd = 2 M f . Rd + 2 B Rd n m + n
Eventually, the third case was concerned with bolt failure, in which connected plates were fully separated. In this case, the design resistance of the T-stub is given by Equation (7):
F 3 . Rd = 2 B Rd
To obtain a ductile behavior, the T-stubs must be designed according to mode 1: flange yielding and higher strength should be achieved for the other failure modes, including bolt failure.
Table 7 denotes the failure modes for all 19 models. For instance, the connection did not have any failure modes by equipping the T-stub section or column bolt with SMA material. Mode 3 failure occurred in most developed models. For example, the column bolts failed by an increase in L2 value of about 50%, although mode 2 failure occurred in the SP04 and SP11 models, having different numbers of beam bolts and flange thicknesses of the T-stub section, respectively.
Based on Figure 9, when applying the lateral loads on the SP18 model, mode 3 developed. It showed critical Von Mises stress in the bolt, which illustrated the failure of the bolt. In addition, the behavior of the T-stub flange remained within a limit of the linear region and did not yield.

5. Effects of SMA on Stiffener, T-Stub, and Bolt Column

An SMA material is recognized as an alloy that can be deformed when the temperature drops, while attaining its initial mode shape without residual displacement after increasing the temperature. Such a feature offers intelligent material behavior when subjected to seismic loads. The effectiveness of the SMA property on the response of models was investigated, in which the T-stubs were equipped with different types of stiffeners.
The newly developed models were named the SP20 to SP43 models, as shown in Table 8. The specifications of the models changed in thickness, number, and type of stiffener and material of components, including the stiffeners, T-stub sections, and column bolts in bolted steel connections. For instance, in the SP28 model, T-stubs were equipped with stiffeners (e.g., type B), and in addition, the material of column bolt was changed from steel to SMA, while as shown in the SP29 model, the T-stubs were equipped with stiffeners (e.g., type A), and the SMA property was assigned to the stiffeners. The stiffeners’ dimensions were determined using ABAQUS optimization scripting. Different dimensions were defined in the ABAQUS scripts, and it updated automatically and compared specific results. Figure 10 shows the details.
The stiffeners are reported to have caused rapid growth in the DST connections’ initial stiffness and maximum force. Their target in this research was to postpone the early yield and increase the tolerance threshold, which occurred by the enlargement of the initial stiffness of the stiffeners’ equipped specimens. Another aim of this section was to modify the stiffeners geometry; there were two types of stiffeners for further investigation.

5.1. Monitoring Behavioural Curve

The force–displacement curves for the SP20 to SP43 models are presented in Figure 11. These curves exhibited the simultaneous effect of the various parameters of importance on the cyclic behavior of bolted steel connection. As a significant result, force values that were enhanced in the models could be observed clearly when comparing with the SP14 and SP15 models for 4% drift. This increasing trend started at approximately 1% drift and was substantially visible between 2% and 4% drifts. In addition, the SP20 to SP43 models indicated a remarkable increase in the residual displacement value compared with the SP15 model. For example, the residual displacement value of the SP30 model increased by 63% when compared with the SP15 model.
As stated in Table 9, an SMA material had beneficial effects on the total performance of DST connections. For example, all three studied parameters (Ki, Kr, and Fmax) for SP26 were higher than those of SP02. This validated the usage of stiffeners, either steel or SMA materials. It consistently increased the initial stiffness (Ki) and ultimate strength (Fmax) of the connection (each by 113.52% and 15%), while the first yield point of the connection was delayed (higher post-yield or residual stiffness).
Another vital element alongside the stiffener was the T-stub section. The overall performance and response of the DST connections depended on the deformation quality and status of the T-stub section. SP15 was an example where it had an SMA type T-stub section, in which the initial slope of the force–displacement curve (Ki) and Fmax experienced a percentage decrease of 27.4% and 11.8%. This reduction showed that SP15 was more flexible and ductile when compared with the original specimen (SP02), whereas the T-stub section could rebound or return to its preliminary position due to the inherent SMA material specifications. These results conclusively supported the reasons for using SMA material.
As shown in Figure 12, for the DST connection, the existence of the SMA column bolt and changes in the number and thickness of the stiffener increased the value of Mmax when compared with SP14, from 19% to 30%. Equipping the connection with the SMA T-stub section and changing the number and thickness of stiffeners led to changes in the Mmax value. This increase was between 19.6% to 34.6% when compared with the SP15 model. As shown in Figure 13, for non-reference models where the column bolt or T-stub section was equipped with SMA property, a change was seen in the number or thickness of the stiffeners simultaneously. The ED capacity of the SP24 and SP25 models increased when compared with the SP14 and SP15 models.

5.2. Equivalent Viscous Damping Ratio

As for further investigation of the connection behavior and response, another index was introduced to show the dissipation potential of the DST connections, exclusively in the specimens equipped with T-stub stiffeners. This index was called the equivalent viscous damping ratio (EVDR), a dimensionless index that could be calculated by Equation (8) [117,118,119]:
ξ = E D 4 π   E So
where ED is the total energy loss per loading cycle and Es is the strain energy in an elastic system with identical maximum deformation and force to a nonlinear system.
Considering Figure 14, Ed and Es are the total energy loss per loading cycle and strain energy in an elastic system with identical maximum deformation and force, respectively. It is clear that Ed for each specimen would increase by drift increment. The findings in this paper provide additional information about the definition of the equivalent viscous damping ratio and its usage in assessing steel or composite connection [117,118,120,121,122].
According to the EVDR/drift figures, the application of stiffeners affected the energy dissipation of the DST connections; EVDR increased up to 20% in a 4% drift. That showed a successful modified and developed edition of an original specimen (a stiffener-free model). Figure 15 shows the subsequent results on the EVDR versus drift level for all stiffened specimens.

5.2.1. Ductility Capacity and Failure Modes

Ductility is the ability to endure the nonlinear deformations without failing. Using this concept in analysis and design can guarantee the optimized performance of a structure, when facing strong lateral displacements. The ductility ratios of the models were calculated by the equivalent energy elastic–plastic (EEEP) method that has been recommended by ASTM E2126-11 [123]. The ductility ratio is defined as the ratio of the ultimate displacement to the yield displacement (rotation), respectively, and is shown in Equation (9).
μ = Δ u Δ y
The connection behavior had a straight/linear relationship with the value of the ductility ratio. The equivalent energy elastic–plastic method used two enclosed areas by intersecting three points on the curve: namely, yield, maximum, and ultimate point. Figure 16 shows that A1 and A2 were created with intersection of the curve lines with equal areas.
As shown in Table 10, the extents of ductility revealed fluctuations from 2.66 to 3.97. The maximum and minimum values were calculated for the SP29 and SP39 models, respectively.
The results could be observed that an increase in the stiffener thickness reduced the ductility ratio only when the stiffener type was equipped with SMA material. For example, in the SP26 model, there was 11% reduction in the ratio in comparison with the SP20 model. Simultaneously, having an SMA T-stub or SMA column bolt increased the ductility ratio. As for the SP33 and SP39 models, the effectiveness of the SMA T-stub and stiffener thickness in the ductility ratio was observed.
Regarding the effect of the number of stiffeners, the ductility ratio in the SP43 model was compared with SP31. The existence of the SMA column bolt and change of stiffener number from one to three resulted in an 18% increase. However, the exact change in the number of stiffeners and of the addition of the SMA T-stub or SMA stiffener in the bolted steel connection led to reducing the ductility ratio.
Table 10 also illustrates the failure modes for the SP20 to SP43 models. It was shown that mode 3 occurred in the models having stiffeners or T-stub with SMA properties. Figure 17 shows a representative mode 3 failure for SP32. The Von Mises distribution in the bolt was evaluated critically, and the bolt failed. However, the distribution of the AC yield in the T-stub revealed that the flange of the T-stub did not yield. The models that included the column bolt with SMA properties did not experience any failure.

5.2.2. Bolt Force, Strain Energy, and Frictional Energy

Figure 18 illustrates the frictional dissipation capacity for 12 specimens that were subjected to cyclic loading. The capacity of the models for bolts having SMA was compared with other models in which the T-stub material or stiffener material was changed from steel to SMA. For example, at the end of lateral loading, the frictional dissipation of the SP28 model was about 28 KN.m. However, the dissipation of models such as SP26 and SP27 was limited to approximately 12.5 KN.m.
Based on the dissipations, the simultaneous change in the number of stiffeners and the material of the stiffeners, bolts, or T-stubs did not significantly affect the frictional dissipation in the bolted steel connection.
The strain energy was obtained from the strain and stress created in the part or body and the volume change of the body undergoing deformations. The strain energy was generated in the elastic phase of the part so that the body sustained some stress due to loading and underwent local deformation and strain. Total deformation led to altering the volume and structure of the part. The strain energy was directly related to the part’s volume variation rate.
The total strain energy consisted of two elastic and non-elastic (plastic) phases. Thus, the figures show that the connection remained elastic in the drifts below 1%, and the elastic strain energy was formed. However, when the rotation increased from 1% to 4%, the connection tolerated plastic deformations due to increased stresses on the connecting parts.
As discussed before, the reason for the lower total energy dissipation and strain energy of SP27 (176.3 KN.m) was type B stiffeners. Since this type of stiffener had drilled holes, the volume of the part was less than the other stiffener type (without drilled holes or the solid one); hence, higher strain energy is observed from SP30 (197.9 KN.m). The reason for the proposed type B stiffener was its high capacity for total energy dissipation. The induced deformation and stress in the stiffener increased with increasing loading intensity, and this enhancement resulted in the loss of connection internal energy through significant plastic strains as shown in Figure 19.
As shown in Figure 20, at the beginning of loading, the force generated in the bolts of all three models was approximately identical. If the loading intensity or the step time increased, the difference between the generated forces of the bolts was insignificant.
In other words, an increment in the bolt force was due to the increase in the connection rotation. At drifts above 3 to 4%, the bolts underwent high tensile stress, and the bolts could be brought to the yield point threshold and could eventually collapse and rupture.
Among three models, the maximum bolt force was observed for SP30, with SMA T-stub (double split tee section). Such an improvement in the bolt force was due to a large amount of plastic deformation in the T section because of high drifts and with respect to the SMA alloy properties. The T-stub sought to restore all deformations and strains created in the preloading state or its initial state by adding large tensile forces to the bolts. The beam rotation caused the T-stub flange to rotate, and the bolts connected to the column flange prevented that rotation. Therefore, the forces in bolts were increased sharply for each cycle.
Ultimately, the addition of the stiffeners had a considerable influence on the behavior and performance of T-stub connections, ultimate moments, and energy dissipation capacity. The ultimate moment was responsible for decaying the internal energy in the connection and reducing the amount of damage induced. The change allowed for the optimization of the proposed connection, which increased the energy dissipation capacity of the T-stub and resulted in material savings for manufacturing stiffeners. Therefore, the use of type B stiffeners caused more energy loss, and the connection withstood more bending moments before failure.

6. Conclusions

This paper presented the cyclic performance and response of double split tee connections, specifically quantifying the addition of stiffeners. The findings revealed that reducing the number of bolts used in the beam caused a decrease in the connection’s ultimate moment by up to 11%. Increasing the thickness of the T-stub, the yield stress of the T-stub and bolts, and the bolt preloads, however, improved the connection’s moment capacity. Adding a column bolt to the SMA decreased the connection’s ultimate moment due to the SMA’s behavior and the high stress levels of austenite and martensite. Additionally, including triangular stiffeners in the models increased the connection’s stiffness, resulting in an increase in the ultimate moment and moment capacity of the models with 15 mm and 25 mm thickness stiffeners by 16% and 21% times compared with the reference models.
The research found that increasing the number of bolts and the yield stress of the T-stub led to an increase in the failure index and a sudden brittle fracture. However, models with a higher T-stub thickness and bolt preload showed an increased failure index. This indicated that the haunch of the T-stub was more likely to experience a high brittle fracture at a 4% drift. The use of stiffeners on the T-stub improved the performance of the DST connection by decreasing the failure index, pressure index, and triaxiality index. Specifically, the failure index was reduced by 69% when using a 15 mm thick stiffener. The stiffeners improved the behavior of the T-stub and reduced its likelihood of failure, making their use in bolted steel connections recommended. Additionally, using an SMA T-stub in the DST connection improved its role as an energy dissipater and the overall performance of the connection.
The study found that the failure mode of the models was mode 1 when the T-stub, bolts, or stiffeners were made of SMA materials. However, when changes in the bolt pretension, the yield stress of the T-stub, or the thickness of the T-stub were made, the failure mode switched to mode 2. Additionally, reducing the number of bolts from six to four or two led to a switch in failure mode from mode 1 to mode 3.
In terms of the energy dissipation capacity, increasing the number of bolts and the thickness and yield stress of the T-stub increased the ED capacity of the beam. The models with SMA material also showed an increase in energy dissipation. On the other hand, using stiffeners on the T-stub also increased the ED capacity, with those using three stiffeners with 2.5 cm thickness showing a 55% improvement in ED capacity.
Additionally, the study found that the Ki value of connections reinforced by steel and SMA stiffeners was greater than that of the reference model. The Kr variable was, however, reduced by 58% when using an SMA stiffener. The equivalent viscous damping ratio (EVDR) was also increased for the modified and stiffened models with steel and SMA stiffeners when compared with the reference model.
In comparison with the stiffener-free specimens, the specimens with steel/SMA stiffeners exhibited higher income energy, primarily due to inelastic/plastic deformation. Despite slight differences, the DST connections with steel/SMA stiffeners demonstrated increased ED capacities, Ki, Kr, and Fmax when compared with the reference model.
Recommendations for further studies include the following:
  • Investigating the use of different types and thicknesses of stiffeners to determine their impacts on the connection’s performance.
  • Examining the effects of different T-stub thickness, yield strength, and bolt preload to improve the ultimate moment and ED capacity of the connection.
  • Evaluating the behavior and performance of the connection with different types of loading conditions, such as cyclic, static, and dynamic loading.
  • Conducting more comprehensive tests, such as shake table testing, to evaluate the real-world performance of the connection in seismic situations.
  • Performing a detailed cost–benefit analysis of using different materials and designs to optimize the practicality and cost-effectiveness of the connection.
  • Investigating the behavior of the DST connection under different loading conditions such as cyclic loading, static loading, and dynamic loading.

Author Contributions

Conceptualization, A.T. and N.A.; methodology, H.H.; software, N.A. and A.T.; validation, H.H., A.T. and N.A.; formal analysis, G.U.; investigation, N.A.; resources, S.Y.; data curation, A.T.; writing—original draft preparation, N.A.; writing—review and editing, G.U.; visualization, A.T. and S.Y.; supervision, G.U.; project administration, H.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data sharing is not applicable to this article. Data is unavailable due to privacy or ethical restrictions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grecea, D.; Dinu, F.; Dubină, D. Performance criteria for MR steel frames in seismic zones. J. Constr. Steel Res. 2004, 60, 739–749. [Google Scholar] [CrossRef]
  2. Carter, C.J.; Iwankiw, N.R. Improved ductility in seismic steel moment frames with dogbone connections. J. Constr. Steel Res. 1998, 46, 448. [Google Scholar] [CrossRef]
  3. Engelhardt, M.; Winneberger, T.; Zekany, A.; Potyraj, T. Experimental Investigation of Dogbone Moment Connections. Eng. J. 1998, 35, 128–139. [Google Scholar]
  4. Montuori, R.; Nastri, E.; Piluso, V.; Troisi, M. Influence of connection typology on seismic response of MR-Frames with and without ‘set-backs’. Earthq. Eng. Struct. Dyn. 2017, 46, 5–25. [Google Scholar] [CrossRef]
  5. Montuori, R.; Nastri, E.; Piluso, V. Advances in theory of plastic mechanism control: Closed form solution for MR-Frames. Earthq. Eng. Struct. Dyn. 2015, 44, 1035–1054. [Google Scholar] [CrossRef]
  6. Faella, C.; Montuori, R.; Piluso, V.; Rizzano, G. Failure mode control: Economy of semi-rigid frames. In Proceedings of the XI European Conference on Earthquake Engineering, Paris, France, 6–11 September 1998. [Google Scholar]
  7. Tenchini, A.; D’Aniello, M.; Rebelo, C.; Landolfo, R.; da Silva, L.S.; Lima, L. Seismic performance of dual-steel moment resisting frames. J. Constr. Steel Res. 2014, 101, 437–454. [Google Scholar] [CrossRef]
  8. Cassiano1a, D.; D’Aniello, M.; Rebelo, C.; Landolfo, R.; da Silva, L.S. Influence of seismic design rules on the robustness of steel moment resisting frames. Steel Compos. Struct. 2016, 21, 479–500. [Google Scholar] [CrossRef]
  9. Paella, C.; Piluso, V.; Rizzano, G. Structural Steel Semirigid Connections: Theory, Design, and Software; CRC Press: Boca Raton, FL, USA, 2000. [Google Scholar]
  10. Kong, Z.; Kim, S.-E. Numerical estimation for initial stiffness and ultimate moment of T-stub connections. J. Constr. Steel Res. 2018, 141, 118–131. [Google Scholar] [CrossRef]
  11. Guo, H.; Liang, G.; Li, Y.; Liu, Y. Q690 high strength steel T-stub tensile behavior: Experimental research and theoretical analysis. J. Constr. Steel Res. 2017, 139, 473–483. [Google Scholar] [CrossRef]
  12. Liang, G.; Guo, H.; Liu, Y.; Li, Y. Q690 high strength steel T-stub tensile behavior: Experimental and numerical analysis. Thin-Walled Struct. 2018, 122, 554–571. [Google Scholar] [CrossRef]
  13. Atasoy, M. Determination of Prying Load on Bolted Connections. Master’s Thesis, Middle East Technical University, Ankara, Türkiye, 2012. [Google Scholar]
  14. Kulak, G.; Fisher, J.; Struik, J. Guide to Design Criteria for Bolted and Riveted Joints, 2nd ed.; American Institute of Steel Construction: Chicago, IL, USA, 2001. [Google Scholar]
  15. Bezerra, L.M.; Bonilla, J.; Freitas, C.S.; Massicotte, B. Behavior of T-stub steel connections bolted to rigid bases. J. Constr. Steel Res. 2022, 192, 107242. [Google Scholar] [CrossRef]
  16. American Institute of Steel Construction. Steel Construction Manual, 14th ed.; American Institute of Steel Construction: Chicago, IL, USA, 1927. [Google Scholar]
  17. Douty, R.; McGuire, W. High strength bolted moment connections. J. Struct. Div. 1965, 91, 101–128. [Google Scholar] [CrossRef]
  18. Nair, R.; Birkemoe, P.; Munse, W. High strength bolts subject to tension and prying. J. Struct. Div. 1974, 100, 351–372. [Google Scholar] [CrossRef]
  19. Zhu, X.; Wu, Z. Study on New Bolted T-Stub Connection with Inserted Plates under Axial and Cyclic Loads. Appl. Sci. 2020, 10, 2631. [Google Scholar] [CrossRef] [Green Version]
  20. ABNT NBR 8800; Design of Steel and Composite Structures for Buildings. ABNT: Rio de Janeiro, Brazil, 2008.
  21. Struik, J.; de Back, J. Tests on Bolted T-Stubs with Respect to a Bolted Beam-to-Column Connections; Stevin Laboratory, Delft University of Technology: Delft, The Netherlands, 1969; pp. 6–69. [Google Scholar]
  22. American Institute of Steel Construction. Load and Resistance Factor Design (LRFD) Specifications for Structural Steel Buildings; American Institute of Steel Construction: Chicago, IL, USA, 1986. [Google Scholar]
  23. American Society of Civil Engineers. Plastic Design in Steel: A Guide and Commentary; Manual and Report of Engineering Practice Ser. No. 41; American Society of Civil Engineers: New York, NY, USA, 1971. [Google Scholar]
  24. American Institute of Steel Construction. Manual of Steel Construction, 7th ed.; American Institute of Steel Construction: Chicago, IL, USA, 1970. [Google Scholar]
  25. European Committee for Standardization (CEN). Eurocode 3: Design of Steel Structures—Part 1-1: General Rules and Rules for Buildings; European Committee for Standardization (CEN): Brussels, Belgium, 1992. [Google Scholar]
  26. Kato, B.; McGuire, W. Analysis of T-stub flange-to-column connections. J. Struct. Div. 1973, 99, 865–888. [Google Scholar] [CrossRef]
  27. Packer, J.; LJ, M. A limit state design method for the tension region of bolted beam-column connections. Struct. Eng. 1977, 55, 446–458. [Google Scholar]
  28. Agerskov, H. Analysis of bolted connections subject to prying. J. Struct. Div. 1977, 103, 2145–2163. [Google Scholar] [CrossRef]
  29. Yee, Y.; Melchers, R. Moment-rotation curves for bolted connections. J. Struct. Eng. 1986, 112, 615–635. [Google Scholar] [CrossRef]
  30. Astaneh, A. Procedure for design and analysis of hanger-type connections. Eng. J. AISC 1985, 22, 63–66. [Google Scholar]
  31. Thornton, W.A. Prying action-a general treatment. Eng. J. 1985, 22, 67–75. [Google Scholar]
  32. Bursi, O.S.; Jaspart, J.P. Calibration of a finite element model for isolated bolted end-plate steel connections. J. Constr. Steel Res. 1997, 44, 225–262. [Google Scholar] [CrossRef]
  33. Swanson, J.A.; Kokan, D.S.; Leon, R.T. Advanced finite element modeling of bolted T-stub connection components. J. Constr. Steel Res. 2002, 58, 1015–1031. [Google Scholar] [CrossRef]
  34. Girão Coelho, A.M.; Bijlaard, F.S.K.; Simões da Silva, L. Experimental assessment of the ductility of extended end plate connections. Eng. Struct. 2004, 26, 1185–1206. [Google Scholar] [CrossRef]
  35. Girão Coelho, A.M. Rotation capacity of partial strength steel joints with three-dimensional finite element approach. Comput. Struct. 2013, 116, 88–97. [Google Scholar] [CrossRef]
  36. Abidelah, A.; Bouchaïr, A.; Kerdal, D.E. Influence of the flexural rigidity of the bolt on the behavior of the T-stub steel connection. Eng. Struct. 2014, 81, 181–194. [Google Scholar] [CrossRef]
  37. Barata, P.; Ribeiro, J.; Rigueiro, C.; Santiago, A.; Rodrigues, J.P. Assessment of the T-stub joint component at ambient and elevated temperatures. Fire Saf. J. 2014, 70, 1–13. [Google Scholar] [CrossRef]
  38. Elflah, M.; Theofanous, M.; Dirar, S. Behaviour of stainless steel beam-to-column joints-part 2: Numerical modelling and parametric study. J. Constr. Steel Res. 2019, 152, 194–212. [Google Scholar] [CrossRef] [Green Version]
  39. Rostami, A.; Moghadam, A.S.; Hosseini, M.; Asghari, N. Evaluation of Formation of Plastic Hinge and Seismic Behavior of Steel Structures Due to Soil–Structure–Tunnel Interaction. J. Earthq. Tsunami 2020, 14, 2050014. [Google Scholar] [CrossRef]
  40. Girão Coelho, A.M.; Bijlaard, F.S.K.; Gresnigt, N.; Simões da Silva, L. Experimental assessment of the behaviour of bolted T-stub connections made up of welded plates. J. Constr. Steel Res. 2004, 60, 269–311. [Google Scholar] [CrossRef]
  41. Gil, B.; Goñi, R.; Bayo, E. Initial stiffness and strength characterization of minor axis T-stub under out-of-plane bending. J. Constr. Steel Res. 2018, 140, 208–221. [Google Scholar] [CrossRef]
  42. Sumner, E.; Mays, T.; Murray, T. Cyclic Testing of Bolted Moment End-Plate Connections; Research Report SAC/BD-00/21. CE/VPI-ST 00/03; Department of Civil and Environmental Engineering, Virginia Polytechnic Institute and State University: Blacksburg, VA, USA, 2000. [Google Scholar]
  43. Gerami, M.; Saberi, H.; Saberi, V.; Saedi Daryan, A. Cyclic behavior of bolted connections with different arrangement of bolts. J. Constr. Steel Res. 2011, 67, 690–705. [Google Scholar] [CrossRef]
  44. Özkılıç, Y.O. Cyclic and monotonic performance of stiffened extended end-plate connections with large-sized bolts and thin end-plates. Bull. Earthq. Eng. 2022, 20, 7441–7475. [Google Scholar] [CrossRef]
  45. Özkılıç, Y.O.; Topkaya, C. Extended end-plate connections for replaceable shear links. Eng. Struct. 2021, 240, 112385. [Google Scholar] [CrossRef]
  46. Özkılıç, Y.O. A comparative study on yield line mechanisms for four bolted extended end plated connection. J. Struct. Mech. 2021, 7, 93–106. [Google Scholar] [CrossRef]
  47. Özkılıç, Y.O. Investigation of the effects of bolt diameter and end-plate thickness on the capacity and failure modes of end-plated beam-to-column connections. Res. Eng. Struct. Mater. 2021, 7, 445–463. [Google Scholar] [CrossRef]
  48. Mistakidis, E.S.; Baniotopoulos, C.C.; Bisbos, C.D.; Panagiotopoulos, P.D. Steel T-Stub connections under static loading: An effective 2-D numerical model. J. Constr. Steel Res. 1997, 44, 51–67. [Google Scholar] [CrossRef]
  49. European Commission. Numerical Aspects for the Simulation of End Plate Connections by the Finite Element Method; European Commission: Brussels, Belgium, 1999; pp. 13–31. [Google Scholar]
  50. Sherbourne, A.N.; Bahaari, M.R. 3D simulation of bolted connections to unstiffened columns—I. T-stub connections. J. Constr. Steel Res. 1996, 40, 169–187. [Google Scholar] [CrossRef]
  51. Bursi, O.S.; Jaspart, J.-P. A refined finite element model for tee-stub and endplate steel connections. In Proceedings of the Giornate Italiane della Costruzione in Acciaio, Riva del Garda, Italy, 15–18 October 1995. [Google Scholar]
  52. Bursi, O.S.; Jaspart, J. Finite element based models for the analysis of bolted beam-to-column steel connections. In Proceedings of the IABSE Colloquium on Semi-Rigid Structural Connections, Istanbul, Turkey, 25–27 September 1996. [Google Scholar]
  53. Bursi, O.S.; Jaspart, J.P. Basic issues in the finite element simulation of extended end plate connections. Comput. Struct. 1998, 69, 361–382. [Google Scholar] [CrossRef]
  54. Bursi, O.S.; Jaspart, J.P. Benchmarks for finite element modelling of bolted steel connections. J. Constr. Steel Res. 1997, 43, 17–42. [Google Scholar] [CrossRef]
  55. Hantouche, E.; Kukreti, A.; Rassati, G.; Swanson, J. Prying models for strength in thick-flange built-up T-stubs with complete joint penetration and fillet welds. J. Struct. Eng. 2015, 141, 04014102. [Google Scholar] [CrossRef]
  56. Hantouche, E.; Kukreti, A.R.; Rassati, G.A. Investigation of secondary prying in thick built-up T-stub connections using nonlinear finite element modeling. Eng. Struct. 2012, 36, 113–122. [Google Scholar] [CrossRef]
  57. Hantouche, E.; Kukreti, A.; Rassati, G.; Swanson, J. Modified stiffness model for thick flange in built-up T-stub connections. J. Constr. Steel Res. 2013, 81, 76–85. [Google Scholar] [CrossRef]
  58. Hantouche, E.; Abboud, N.H. Stiffness modeling of bolted thick built-up T-stub connections including secondary prying effect. J. Constr. Steel Res. 2014, 95, 279–289. [Google Scholar] [CrossRef]
  59. Hantouche, E.; Rassati, G.; Kukreti, A.; Swanson, J. Built-up T-stub connections for moment resisting frames: Experimental and finite element investigation for prequalification. Eng. Struct. 2012, 43, 139–148. [Google Scholar] [CrossRef]
  60. Lemonis, M.; Gantes, C. Finite Element Modeling of T-Stub Steel Connections. In Proceedings of the 4th GRACM Congress on Computational Mechanics GRACM, Patras, Greece, 27–29 June 2002. [Google Scholar]
  61. Lemonis, M.; Gantes, C. Numerical analysis of simple and preloaded T-stub steel connections. In Computational Fluid and Solid Mechanics 2003; Bathe, K.J., Ed.; Elsevier Science Ltd.: Oxford, UK, 2003; pp. 414–417. [Google Scholar] [CrossRef]
  62. Gantes, C.J.; Lemonis, M.E. Influence of equivalent bolt length in finite element modeling of T-stub steel connections. Comput. Struct. 2003, 81, 595–604. [Google Scholar] [CrossRef]
  63. Lemonis, M.; Gantes, C. Incremental modeling of T-stub connections. J. Mech. Mater. Struct. 2006, 1, 1135–1159. [Google Scholar] [CrossRef] [Green Version]
  64. Soltani, R.; Kerdal, D. Behaviour of elementary bolted steel T-stub connections: An evaluation of EC3 design procedure. Turk. J. Eng. Environ. Sci. 2011, 35, 1–20. [Google Scholar]
  65. Ceniceros, J.F. Numerical-Informational Methodology for Characterising Steel Bolted Components Coupling Finite Element Simulations and Soft Computing Techniques; Universidad de La Rioja: Logroño, Spain, 2015. [Google Scholar]
  66. Fernandez-Ceniceros, J.; Sanz-Garcia, A.; Antoñanzas-Torres, F.; Martinez-de-Pison, F.J. A numerical-informational approach for characterising the ductile behaviour of the T-stub component. Part 1: Refined finite element model and test validation. Eng. Struct. 2015, 82, 236–248. [Google Scholar] [CrossRef]
  67. Fernandez-Ceniceros, J.; Sanz-Garcia, A.; Antoñanzas-Torres, F.; Martinez-de-Pison, F.J. A numerical-informational approach for characterising the ductile behaviour of the T-stub component. Part 2: Parsimonious soft-computing-based metamodel. Eng. Struct. 2015, 82, 249–260. [Google Scholar] [CrossRef]
  68. Kombate, T.J.D.L.C.; Taşkın, K. State-of-the-art review on the behaviour of T-stubs and prying action. J. Constr. Steel Res. 2022, 191, 107203. [Google Scholar] [CrossRef]
  69. Bravo, M.A.; Herrera, R.A. Performance under cyclic load of built-up T-stubs for Double T moment connections. J. Constr. Steel Res. 2014, 103, 117–130. [Google Scholar] [CrossRef]
  70. Herrera, R.A.; Bravo, M.; Gómez, G.; Aedo, G. Performance of built-up T-stubs for Double T moment connections. J. Constr. Steel Res. 2013, 88, 289–295. [Google Scholar] [CrossRef]
  71. Vahdati, V.J.; Yaghoubi, S.; Torabipour, A.; Correia, J.A.F.O.; Fazeres-Ferradosa, T.; Taveira-Pinto, F. Combined solutions to reduce scour around complex foundations: An experimental study. Mar. Syst. Ocean Technol. 2020, 15, 81–93. [Google Scholar] [CrossRef]
  72. Ribeiro, J.; Santiago, A.; Rigueiro, C.; Barata, P.; Veljkovic, M. Numerical assessment of T-stub component subjected to impact loading. Eng. Struct. 2016, 106, 450–460. [Google Scholar] [CrossRef]
  73. Özkılıç, Y.O.; Topkaya, C. The plastic and the ultimate resistance of four-bolt extended end-plate connections. J. Constr. Steel Res. 2021, 181, 106614. [Google Scholar] [CrossRef]
  74. Özkılıç, Y.O. The effects of stiffener configuration on stiffened T-stubs. Steel Compos. Struct. 2022, 44, 475–488. [Google Scholar]
  75. Özkılıç, Y.O. The capacities of unstiffened T-stubs with thin plates and large bolts. J. Constr. Steel Res. 2021, 186, 106908. [Google Scholar] [CrossRef]
  76. Özkılıç, Y.O. The capacities of thin plated stiffened T-stubs. J. Constr. Steel Res. 2021, 186, 106912. [Google Scholar] [CrossRef]
  77. Daryan, A.S.; Sadri, M.; Saberi, H.; Saberi, V.; Moghadas, A.B. Behavior of semi-rigid connections and semi-rigid frames. Struct. Des. Tall Spec. Build. 2014, 23, 210–238. [Google Scholar] [CrossRef]
  78. Shiravand, M.R.; Torabipour, A.; Mahboubi, S. Parametric Study on Effect of Adding Stiffener to Post-tensioned Steel Connection. Int. J. Steel Struct. 2019, 19, 478–494. [Google Scholar] [CrossRef]
  79. Abdollahzadeh, G.; BarehKheil, N.A.; Torabipour, A. Moment–rotation response of post-tensioned self-centring beam–column connection. Proc. Inst. Civ. Eng. Struct. Build. 2020, 173, 440–457. [Google Scholar] [CrossRef]
  80. Torabipour, A.; Abdollahi, A.; Yaghoubi, S. Assess the Effectiveness of SMA on Response of Steel Connection with PT Strands. KSCE J. Civ. Eng. 2019, 23, 5133–5142. [Google Scholar] [CrossRef]
  81. Maghsoudi, A.; Maghsoudi, M.; Haghighi, H. FE Modeling and Experimental Comparison of RC Beams Consisting Shape Memory Alloys under Cyclic Loading. Amirkabir J. Civ. Eng. 2018, 49, 745–754. [Google Scholar] [CrossRef]
  82. Asada, H.; Tanaka, T.; Yamada, S.; Matoba, H. Proposal for seismic retrofit of beam-to-column connection by the addition of H-section haunches to beams using bolt connection. Int. J. Steel Struct. 2014, 14, 865–871. [Google Scholar] [CrossRef]
  83. Sebbagh, H.R.; Kerdal, D.E.D.; Abidelah, A.; Bouchaïr, A. T-stubs with two and four bolts under monotonic and cyclic loading. J. Constr. Steel Res. 2021, 178, 106486. [Google Scholar] [CrossRef]
  84. Yadupriya, M.C.; Rajeevan, B. Numerical analysis on bolted T-stub to square hollow section connections under tension. 2022. Available online: https://ssrn.com/abstract=4269427 (accessed on 6 November 2022).
  85. Tartaglia, R.; D’Aniello, M.; Zimbru, M. Experimental and numerical study on the T-Stub behaviour with preloaded bolts under large deformations. Structures 2020, 27, 2137–2155. [Google Scholar] [CrossRef]
  86. Wang, Y.; Qiao, X.; Jia, L.; Zhang, T.; Jiang, Q. Experimental research on seismic behavior of stainless steel beam-column connections with different link modes. J. Southeast Univ. 2018, 48, 316–322. [Google Scholar]
  87. Liu, M.; Shi, G. Cyclic loading tests of duplex stainless steel beam-to-column joints with WUF-W connections. In Proceedings of the Ninth International Conference on Advances in Steel Structures (ICASS), Hong Kong, China, 5–7 December 2018; pp. 947–956. [Google Scholar]
  88. Bu, Y.; Wang, Y.; Zhao, Y. Study of stainless steel bolted extended end-plate joints under seismic loading. Thin-Walled Struct. 2019, 144, 106255. [Google Scholar] [CrossRef]
  89. Yuan, H.X.; Liu, X.H.; Liu, J.L.; Theofanous, M. Cyclic behaviour and hysteretic model of austenitic stainless steel bolted T-stubs. J. Constr. Steel Res. 2021, 182, 106659. [Google Scholar] [CrossRef]
  90. Zhao, M.S.; Lee, C.K.; Chiew, S.P. Tensile behavior of high performance structural steel T-stub joints. J. Constr. Steel Res. 2016, 122, 316–325. [Google Scholar] [CrossRef] [Green Version]
  91. Sun, F.F.; Xue, X.Y.; Xiao, Y.; Le, Y.M.; Li, G.Q. Effect of welding and complex loads on the high-strength steel T-stub connection. J. Constr. Steel Res. 2018, 150, 76–86. [Google Scholar] [CrossRef]
  92. Wang, Z.Y.; Tizani, W.; Wang, Q.Y. Strength and initial stiffness of a blind-bolt connection based on the T-stub model. Eng. Struct. 2010, 32, 2505–2517. [Google Scholar] [CrossRef]
  93. Yang, J.-G.; Lee, S.-K.; Moon, J.-K.; Baek, M.-C. Prediction models for the initial rotational stiffness and ultimate plastic moment for a double split tee connection. Int. J. Steel Struct. 2014, 14, 213–222. [Google Scholar] [CrossRef]
  94. SAC Joint Venture. Protocol for Fabrication, Inspection, Testing, and Documentation of Beam-Column Connection Tests and Other Experimental Specimens; Rep. No. SAC/BD-97; SAC Joint Venture: Redwood City, CA, USA, 1997; p. 2. [Google Scholar]
  95. Liu, Y.; Málaga-Chuquitaype, C.; Elghazouli, A.Y. Response and component characterisation of semi-rigid connections to tubular columns under axial loads. Eng. Struct. 2012, 41, 510–532. [Google Scholar] [CrossRef] [Green Version]
  96. Massimo, L.; Gianvittorio, R.; Aldina, S.; da Silva Luis, S. Experimental analysis and mechanical modeling of T-stubs with four bolts per row. J. Constr. Steel Res. 2014, 101, 158–174. [Google Scholar] [CrossRef]
  97. De Matteis, G.; Mandara, A.; Mazzolani, F.M. T-stub aluminium joints: Influence of behavioural parameters. Comput. Struct. 2000, 78, 311–327. [Google Scholar] [CrossRef]
  98. Wang, Z.Y.; Wang, Q.-Y. Yield and ultimate strengths determination of a blind bolted endplate connection to square hollow section column. Eng. Struct. 2016, 111, 345–369. [Google Scholar] [CrossRef]
  99. Yapici, O.; Theofanous, M.; Yuan, H.; Dirar, S.; Afshan, S. Numerical simulation and design of ferritic stainless steel bolted T-stubs in tension. J. Constr. Steel Res. 2022, 198, 107555. [Google Scholar] [CrossRef]
  100. Royal Civil Service Commission. Specification for Structural Joints Using High-Strength Bolts; American Institute of Steel Construction: Chicago, IL, USA, 2014. [Google Scholar]
  101. American Institute of Steel Construction. Specification for Structural Steel Buildings. In An American National Standard; American Institute of Steel Construction: Chicago, IL, USA, 2016. [Google Scholar]
  102. DesRoches, R.; McCormick, J.; Delemont, M. Cyclic properties of superelastic shape memory alloy wires and bars. J. Struct. Eng. 2004, 130, 38–46. [Google Scholar] [CrossRef]
  103. Wang, B.; Zhu, S.; Qiu, C.-X.; Jin, H. High-performance self-centering steel columns with shape memory alloy bolts: Design procedure and experimental evaluation. Eng. Struct. 2019, 182, 446–458. [Google Scholar] [CrossRef]
  104. Wang, B.; Zhu, S. Cyclic tension–compression behavior of superelastic shape memory alloy bars with buckling-restrained devices. Constr. Build. Mater. 2018, 186, 103–113. [Google Scholar] [CrossRef]
  105. Pereiro-Barceló, J.; Bonet, J.L.; Cabañero-Escudero, B.; Martínez-Jaén, B. Cyclic behavior of hybrid RC columns using High-Performance Fiber-Reinforced Concrete and Ni-Ti SMA bars in critical regions. Compos. Struct. 2019, 212, 207–219. [Google Scholar] [CrossRef]
  106. Chowdhury, M.A.; Rahmzadeh, A.; Moradi, S.; Alam, M.S. Feasibility of using reduced length superelastic shape memory alloy strands in post-tensioned steel beam–column connections. J. Intell. Mater. Syst. Struct. 2019, 30, 283–307. [Google Scholar] [CrossRef]
  107. Özkılıç, Y.O. A new replaceable fuse for moment resisting frames: Replaceable bolted reduced beam section connections. Steel Compos. Struct. 2020, 35, 353–370. [Google Scholar] [CrossRef]
  108. Ricles, J.M.; Sause, R.; Garlock, M.M.; Zhao, C. Posttensioned Seismic-Resistant Connections for Steel Frames. J. Struct. Eng. 2001, 127, 113–121. [Google Scholar] [CrossRef]
  109. Bozkurt, M.B.; Kazemzadeh Azad, S.; Topkaya, C. Development of detachable replaceable links for eccentrically braced frames. Earthq. Eng. Struct. Dyn. 2019, 48, 1134–1155. [Google Scholar] [CrossRef]
  110. Bozkurt, M.B.; Topkaya, C. Replaceable links with direct brace attachments for eccentrically braced frames. Earthq. Eng. Struct. Dyn. 2017, 46, 2121–2139. [Google Scholar] [CrossRef]
  111. Bozkurt, M.B.; Topkaya, C. Replaceable links with gusseted brace joints for eccentrically braced frames. Soil Dyn. Earthq. Eng. 2018, 115, 305–318. [Google Scholar] [CrossRef]
  112. Özkılıç, Y.O. Optimized stiffener detailing for shear links in eccentrically braced frames. Steel Compos. Struct. Int. J. 2021, 39, 35–50. [Google Scholar] [CrossRef]
  113. Özkılıç, Y.O.; Bozkurt, M.B.; Topkaya, C. Mid-spliced end-plated replaceable links for eccentrically braced frames. Eng. Struct. 2021, 237, 112225. [Google Scholar] [CrossRef]
  114. Özkılıç, Y.O. Interaction of flange and web slenderness, overstrength factor and proposed stiffener arrangements for long links. J. Constr. Steel Res. 2022, 190, 107150. [Google Scholar] [CrossRef]
  115. Korkmaz, H.H.; Dere, Y.; Özkılıç, Y.O.; Bozkurt, M.B.; Ecemiş, A.S.; Özdoner, N. Excessive snow induced steel roof failures in Turkey. Eng. Fail. Anal. 2022, 141, 106661. [Google Scholar] [CrossRef]
  116. El-Tawil, S.; Mikesell, T.; Vidarsson, E.; Kunnath, S. Strength and Ductility of FR Welded-Bolted Connections; Department of Civil and Environmental Engineering, University of Central Florida: Orlando, FL, USA, 1998; p. 150. [Google Scholar]
  117. Fang, C.; Wang, W.; He, C.; Chen, Y. Self-centring behaviour of steel and steel-concrete composite connections equipped with NiTi SMA bolts. Eng. Struct. 2017, 150, 390–408. [Google Scholar] [CrossRef]
  118. Fang, C.; Yam, M.C.H.; Lam, A.C.C.; Xie, L. Cyclic performance of extended end-plate connections equipped with shape memory alloy bolts. J. Constr. Steel Res. 2014, 94, 122–136. [Google Scholar] [CrossRef]
  119. Chopra, A.K. Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th ed.; Prentice Hall: Englewood Cliffs, NJ, USA, 2011. [Google Scholar]
  120. He, M.-j.; Liu, H.-f. Comparison of glulam post-to-beam connections reinforced by two different dowel-type fasteners. Constr. Build. Mater. 2015, 99, 99–108. [Google Scholar] [CrossRef]
  121. Elmenshawi, A.; Brown, T. Hysteretic energy and damping capacity of flexural elements constructed with different concrete strengths. Eng. Struct. 2010, 32, 297–305. [Google Scholar] [CrossRef]
  122. Speicher, M.S.; DesRoches, R.; Leon, R.T. Experimental results of a NiTi shape memory alloy (SMA)-based recentering beam-column connection. Eng. Struct. 2011, 33, 2448–2457. [Google Scholar] [CrossRef]
  123. American Society for Testing and Materials. Standard Test Methods for Cyclic (Reversed) Load Test for Shear Resistance of Vertical Elements of the Lateral Force Resisting Systems for Buildings; American Society for Testing and Materials: West Conshohocken, PA, USA, 2019; p. 15. [Google Scholar]
Figure 1. Bolted steel T-stub connection schematic and deformation of T-stub.
Figure 1. Bolted steel T-stub connection schematic and deformation of T-stub.
Civileng 04 00008 g001
Figure 2. MRF under cyclic loading. (a) The MRF deflection diagram of the MRF, (b) moment diagram, (c) exterior connection, and (d) installed substructure in numerical modeling, reproduced with permission of Gerami et al. [43].
Figure 2. MRF under cyclic loading. (a) The MRF deflection diagram of the MRF, (b) moment diagram, (c) exterior connection, and (d) installed substructure in numerical modeling, reproduced with permission of Gerami et al. [43].
Civileng 04 00008 g002
Figure 3. Details of T-stub connection mesh: (a) overview, (b) bolt, and (c) T-stub.
Figure 3. Details of T-stub connection mesh: (a) overview, (b) bolt, and (c) T-stub.
Civileng 04 00008 g003
Figure 4. (a) Hysteresis curve and (b) Von Mises stress distribution for G310-T19-B400 model.
Figure 4. (a) Hysteresis curve and (b) Von Mises stress distribution for G310-T19-B400 model.
Civileng 04 00008 g004
Figure 5. (a) The arrangement of bolt holes on connected T-stub to beam. (b) General schematic of different parameters in T-stub profile.
Figure 5. (a) The arrangement of bolt holes on connected T-stub to beam. (b) General schematic of different parameters in T-stub profile.
Civileng 04 00008 g005
Figure 6. Response variables.
Figure 6. Response variables.
Civileng 04 00008 g006
Figure 7. (a) Strain energy-displacement curve. (b). Force- displacement curve (c) Von Mises distribution.
Figure 7. (a) Strain energy-displacement curve. (b). Force- displacement curve (c) Von Mises distribution.
Civileng 04 00008 g007
Figure 8. Strain–stress distribution of SMA T-stub under cyclic loading.
Figure 8. Strain–stress distribution of SMA T-stub under cyclic loading.
Civileng 04 00008 g008
Figure 9. Stress concentration on SP18 model. (a) General view, (b) bolt, and (c) T-stub under cyclic loading.
Figure 9. Stress concentration on SP18 model. (a) General view, (b) bolt, and (c) T-stub under cyclic loading.
Civileng 04 00008 g009
Figure 10. (A,B) Geometric dimensioning of stiffener types.
Figure 10. (A,B) Geometric dimensioning of stiffener types.
Civileng 04 00008 g010
Figure 11. Force–displacement curves.
Figure 11. Force–displacement curves.
Civileng 04 00008 g011
Figure 12. Mmax of all connections.
Figure 12. Mmax of all connections.
Civileng 04 00008 g012
Figure 13. ED capacity of all connections.
Figure 13. ED capacity of all connections.
Civileng 04 00008 g013
Figure 14. Illustration of total energy dissipation E D and maximum strain energy E So [119].
Figure 14. Illustration of total energy dissipation E D and maximum strain energy E So [119].
Civileng 04 00008 g014
Figure 15. EVDR versus drift level.
Figure 15. EVDR versus drift level.
Civileng 04 00008 g015aCivileng 04 00008 g015b
Figure 16. Equivalent energy elastic–plastic (EEEP) method.
Figure 16. Equivalent energy elastic–plastic (EEEP) method.
Civileng 04 00008 g016
Figure 17. Distribution of Von Mises and AC yield (MPa) for the SP32 model.
Figure 17. Distribution of Von Mises and AC yield (MPa) for the SP32 model.
Civileng 04 00008 g017
Figure 18. Frictional dissipation and drift.
Figure 18. Frictional dissipation and drift.
Civileng 04 00008 g018
Figure 19. Strain energy and drift.
Figure 19. Strain energy and drift.
Civileng 04 00008 g019
Figure 20. Bolt force and time.
Figure 20. Bolt force and time.
Civileng 04 00008 g020
Table 1. Value of Fy and Fu.
Table 1. Value of Fy and Fu.
ComponentsFy (MPa)Fu (MPa)
Bolt912.7976.1
T-stub399.61435.85
Beam or column399.61435.85
Table 2. Comparing experimental results with numerical modeling results.
Table 2. Comparing experimental results with numerical modeling results.
Sample ModelM0 (KN.m) θ max (%) Mmax (KN.m)
G260-T15-B350PS1Ex257.620.087451.22
PS1Nu251.880.09442.48
Error percent2.233.451.937
G310-T15-B400PS2Ex197.420.098402.4
PS2Nu195.230.1399.61
Error percent1.112.0410.69
G260-T19-B350PS3Ex411.110.037454.36
PS3Nu408.340.038435.85
Error percent0.6742.74.07
G310-T19-B400PS4Ex329.860.13448.68
PS4Nu324.520.125443.17
Error percent1.623.841.23
G260-T21-B350PS5Ex458.080.023502.84
PS5Nu452.340.024495.51
Error percent1.2534.371.46
G310-T21-B400PS6Ex411.780.0385447.97
PS6Nu407.470.04440.73
Error percent1.053.91.62
Table 3. Developed finite element models.
Table 3. Developed finite element models.
Sample Modeltf
(mm)
Fyt
(T-Stub)
(KN)
Bolt Pretension
(KN)
* No
(Bolt Beam)
Fyb
(Bolt)
(KN)
L1
(mm)
L2
(mm)
Stiffener of
T-Stub
(mm)
Equipped SMA Material
SP0119538.51656912.750150….….
SP0219538.51656912.7100100….….
SP0319538.51656912.715050….….
SP0419538.51652912.7100100….….
SP0519538.51654912.7100100….….
SP0619538.51656950.7100100….….
SP0719538.51656874.7100100….….
SP08196201656912.7100100….….
SP09194581656912.7100100….….
SP1023538.51656912.7100100….….
SP1115538.51656912.7100100….….
SP1219538.51906912.7100100….….
SP1319538.51406912.7100100….….
SP1419538.51656*100100….Bolt column
SP1519*1656912.7100100….T-stub
SP1619538.51656912.71001003PL169 × 15….
SP1719538.51656912.71001003PL169 × 25….
SP1819538.51656912.7100100PL169 × 15….
SP1919538.51656912.7100100PL169 × 25….
* Defines yield strength of SMA T-stub and bolt as discussed in DesRoches et al. [102].
Table 4. Moment capacity, ED capacity, and RD.
Table 4. Moment capacity, ED capacity, and RD.
SpecimenSP01SP02SP03SP04SP05SP06SP07SP08SP09SP10
Mmax(KN.m)455.25450.9427.82399.14447.83453.94448.2455.36444.46465.3
ED(KN.m)142.9138110.287.96136.7139.9135.9140.81135.34140
RD(mm)28.6527.1825.4427.582727.2326.327.6526.8128.9
SpecimenSP11SP12SP13SP14SP15SP16SP17SP18SP19
Mmax(KN.m)440.46452.56449.17417399.73541.81557.5523.47545.2
ED(KN.m)137140.9134.8142.0584.75192.1213.6181.8196.85
RD(mm)26.8328.1826.5825.9218.7327.3828.2526.3327.84
Table 5. Response variables.
Table 5. Response variables.
SpecimenSP01SP02SP03SP04SP05SP06SP07SP08SP09SP10
Ki (KN.m)21,70924,07224,84521,14819,86019,85120,35421,05619,63819,758
Kr (KN.m)10,7127130371012,476591857206729599464105980
Fmax (KN)737.62777.43781.47688.17772.12775.76774.37785.11766.31776.38
SpecimenSP11SP12SP13SP14SP15SP16SP17SP18SP19
Ki (KN.m)20,25520,65420,06212,14717,47134,44048,68429,24235,300
Kr (KN.m)591735006037416844174302374758836589
Fmax (KN)776.65780.27774.43719.13689.2934.16961.2902.54940
Table 6. Failure indexes for the developed models.
Table 6. Failure indexes for the developed models.
Sample Model   σ ¯ σ m PEEQPIMIPlasticity IndexTIRI
SP01694.2−567.4991.09814−1.05381.2891463.1547−0.81753.7428
SP02694.2−497.6910.511864−0.92421.2891215.8853−0.71691.5003
SP03667.68−470.1560.341547−0.87301.2399144.0518−0.70420.9821
SP04694.2−419.1910.241981−0.77841.2891102.0586−0.60380.5986
SP05694.2−468.1810.566258−0.86941.2891238.8266−0.67441.5572
SP06694.2−497.2930.511835−0.92351.2891215.8730−0.71631.4990
SP07676.48−484.3840.508633−0.89951.2561214.5225−0.71611.4889
SP08694.2−509.1960.470578−0.94561.2891198.4723−0.73351.4140
SP09694.2−488.2350.544232−0.90661.2891229.5369−0.70331.5629
SP10694.2−499.250.51937−0.92711.2891219.0510−0.71921.5274
SP11694.2−493.4450.492601−0.91631.2891207.7608−0.71081.4307
SP12694.2−498.050.527099−0.92491.2891222.3108−0.71741.5462
SP13646.8−469.9760.486086−0.87271.2011205.0130−0.72661.4457
SP14694.2−494.3920.472123−0.91811.2891199.124−0.71221.374
SP151309.81−347.0980.025471−0.64452.432310.7427−0.2650.03790
SP16596.152−386.190.07693−0.71711.107132.4463−0.64780.2033
SP17553.96−357.880.030672−0.66461.028712.9363−0.6460.0808
SP18665.99−433.9440.170127−0.80581.236771.7532−0.65160.4521
SP19614.55−420.3170.101485−0.78051.141242.8026−0.68390.2831
Table 7. Failure modes in developed models.
Table 7. Failure modes in developed models.
SpecimenSP01SP02SP03SP04SP05SP06SP07SP08SP09SP10
Failure modeMode 3Mode 3Mode 3Mode 2Mode 3Mode 3N/AMode 3Mode 3Mode 3
SpecimenSP11SP12SP13SP14SP15SP16SP17SP18SP19
Failure modeMode 2Mode 3N/AN/AN/AMode 3Mode 3Mode 3Mode 3
Table 8. Classification and specification of developed models.
Table 8. Classification and specification of developed models.
Sample Modeltf
(mm)
No (Bolt Beam)No (Bolt Column)L1 (mm)Thickness of Stiffener (mm)Number of StiffenerType of StiffenerEquipped SMA Material
SP201964100153BStiffener
SP211964100153BT-stub
SP221964100153BBolt column
SP231964100153AStiffener
SP241964100153AT-stub
SP251964100153ABolt column
SP261964100253BStiffener
SP271964100253BT-stub
SP281964100253BBolt column
SP291964100253AStiffener
SP301964100253AT-stub
SP311964100253ABolt column
SP321964100151BStiffener
SP331964100151BT-stub
SP341964100151BBolt column
SP351964100151AStiffener
SP361964100151AT-stub
SP371964100151ABolt column
SP381964100251BStiffener
SP391964100251BT-stub
SP401964100251BBolt column
SP411964100251AStiffener
SP421964100251AT-stub
SP431964100251ABolt column
Table 9. The parameters of the response curve.
Table 9. The parameters of the response curve.
SpecimenSP20SP21SP22SP23SP24SP25SP26SP27SP28SP29SP30SP31
Ki (KN.m)52,45240,75840,55053,81350,83946,14251,40049,36543417637,81262,67733,782
Kr (KN.m)6024.24966.76976.22995.228605901.160492787.56061.42888.42964.76884.8
Fmax (KN)903.91859.88893.02905.02873.7884.59898.64874898.98936.24927.67933.52
SpecimenSP32SP33SP34SP35SP36SP37SP38SP39SP40SP41SP42SP43
Ki (KN.m)48,16845,03639,20148,72043,26741,32649,44444,39239,72255,15355,66648,044
Kr (KN.m)2613.235712794.33436.83510.328702998.634023082.325722989.22778.7
Fmax (KN)875.46826.82850.24869.38830.87856.18875.46824.27854.91912.73880.57916.7
Table 10. Ductility and failure modes.
Table 10. Ductility and failure modes.
SpecimenSP20SP21SP22SP23SP24SP25SP26SP27SP28SP29SP30SP31
θ y (rad)0.0120.0130.0120.0110.0130.0110.0130.0120.010.0130.0110.011
θ u (rad)0.03490.03490.03480.03450.03490.03440.03490.03490.03490.03480.03490.0349
My (KN.m)385.92373.61382.11385.37376.4376.19382.45375.51384.48399.42392.52395.94
μ = θ u θ y 2.992.682.883.112.673.172.682.783.332.663.043.24
Failure mode33N/A33N/A33N/A33N/A
SpecimenSP32SP33SP34SP35SP36SP37SP38SP39SP40SP41SP42SP43
θ y (rad)0.010.010.0120.010.0090.0130.0110.0080.0120.0110.0090.012
θ u (rad)0.03480.03450.03410.03340.03470.03470.03480.03470.03460.03480.03470.0346
My (KN.m)377.55359.25369.42375.06361.01372377.55358.14371.45391.08382.6387
μ = θ u θ y 3.323.842.813.253.612.683.113.972.853.073.822.74
Failure mode33N/A33N/A33N/A33N/A
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Torabipour, A.; Asghari, N.; Haghighi, H.; Yaghoubi, S.; Urgessa, G. Assessing Effectiveness of Shape Memory Alloys on the Response of Bolted T-Stub Connections Subjected to Cyclic Loading. CivilEng 2023, 4, 105-133. https://doi.org/10.3390/civileng4010008

AMA Style

Torabipour A, Asghari N, Haghighi H, Yaghoubi S, Urgessa G. Assessing Effectiveness of Shape Memory Alloys on the Response of Bolted T-Stub Connections Subjected to Cyclic Loading. CivilEng. 2023; 4(1):105-133. https://doi.org/10.3390/civileng4010008

Chicago/Turabian Style

Torabipour, Ahmadreza, Nima Asghari, Homa Haghighi, Shaghayegh Yaghoubi, and Girum Urgessa. 2023. "Assessing Effectiveness of Shape Memory Alloys on the Response of Bolted T-Stub Connections Subjected to Cyclic Loading" CivilEng 4, no. 1: 105-133. https://doi.org/10.3390/civileng4010008

Article Metrics

Back to TopTop