Open access
Technical Papers
Sep 28, 2022

Artificial Neural Network–Based Predictive Tool for Modeling of Self-Centering Endplate Connections with SMA Bolts

Publication: Journal of Structural Engineering
Volume 148, Issue 12

Abstract

Endplate moment connections with shape memory alloy (SMA) bolts provide self-centering for the seismic resilience of structures. Predicting the self-centering response of these new beam–column connections, which have not been codified yet, requires conducting experimental tests or detailed continuum finite-element simulations. Computationally efficient predictive tools are needed to facilitate the analysis, design, and assessment of self-centering connections and moment frames. In this paper, artificial neural networks (ANNs) are used to develop a MATLAB tool for predicting the moment-rotation backbone and self-centering response of extended endplate connections with SMA bolts. As the input for the neural networks, the predictive model development employs a design database of response parameters from 72 finite-element (FE) models and experimental tests of seven beam–column connection specimens. Neural networks are trained for seven response parameters, and the trained networks are used to develop a graphical user interface (GUI). The coefficient of determination for the trained ANNs is in the range of 0.92 to 0.99, indicating acceptable prediction accuracy. Furthermore, optimization studies using a multiobjective genetic algorithm are performed, seeking the minimization of material use (steel and SMA) and improved connection-response characteristics (i.e., stiffness, strength, and ductility). A phenomenological model of SMA connections is also developed in OpenSees. The use of the ANN-based predictive tool for accurate and efficient modeling of SMA-based connections and self-centering moment-resisting frames is illustrated. The computation time for predicting the moment-rotation response of a typical SMA connection is significantly reduced from seven hours in ANSYS to only three minutes in OpenSees while providing the same level of response prediction accuracy. Furthermore, the optimization results are confirmed by performing nonlinear pushover and response history analyses.

Introduction

Shape memory alloys (SMAs) are a class of metallic alloys that can recover their original shape after undergoing large deformations. The superelasticity or recentering capability of SMAs upon mechanical unloading is due to their material phase transformation. SMAs have found several applications for seismic damage mitigation of civil engineering structures (Wilson and Wesolowsky 2005). Several researchers have previously investigated new applications of SMAs for preventing damage and residual deformation in steel beam-to-column connections. Some examples are connections with SMA tendons (Speicher et al. 2011), SMA ring springs (Wang et al. 2017), superelastic SMA bolts with steel angles (Wang et al. 2017), SMA plates (Moradi and Alam 2015a), and extended endplate connections with superelastic SMA bolts (Fang et al. 2013). Ma et al. (2007, 2008) presented proof-of-concept extended endplate connections in which high-strength bolts were replaced with superelastic SMA bolts (as shown in Fig. 1). Based on the results of their numerical studies, SMA-based extended endplate connections exhibited self-centering while providing a moderate level of energy dissipation capacity. The SMA-based connections proposed by Ma et al. (2007, 2008) were further explored by Fang et al. (2013). They experimentally tested seven SMA-based connection specimens and one conventional extended endplate connection. The test results confirmed that SMA bolts provide self-centering while other components, including endplates, beams, and columns, essentially remain elastic.
Fig. 1. Extended endplate beam–column connection equipped with SMA bolts.
For predicting the response of self-centering endplate connections with superelastic SMA bolts, continuum finite-element (FE) models (e.g., Mohammadi Nia and Moradi 2020; Yam et al. 2015) can be used. However, the development of high-fidelity FE models of connections is computationally expensive. The use of these models for simulating the response of self-centering frames is thus practically impossible. Developing computationally efficient predictive tools is necessary to facilitate the rapid analysis and performance-based design and assessment of SMA-based connections or frames in research and practice. Previously, the authors used a response surface method (RSM) to develop surrogate models for endplate connections with SMA bolts (Mohammadi Nia and Moradi 2021). However, the models were limited to a beam depth of 450 mm to 610 mm. This paper employs an artificial neural network (ANN), which better captures nonlinearity than RSM (Hammoudi et al. 2019; Yaro et al. 2022), to develop a new publicly accessible predictive tool for efficiently modeling SMA-based connections and self-centering frames. The development and verification of this tool are presented, along with the illustration of its use for modeling SMA-based connections and moment frames. A broader range is considered for the beam depth, and experimentally tested specimens are included in the database.
Artificial intelligence (AI), notably machine learning (ML), has been used widely in different areas of structural engineering to solve different problems. Some examples are predicting the resistance of structural members (Caglar 2009; Degtyarev and Naser 2021; Zarringol et al. 2021, among others), fire resistance of structures (e.g., Naser et al. 2021; Xu et al. 2012), structural health monitoring, and damage detection (e.g., Huang and Burton 2019; Sediek et al. 2022).
In traditional statistical regression techniques, a predefined structure of the regression model is required; however, in many cases, this approach is not the optimal approach (Majidifard et al. 2019). ML techniques can derive the relationship between inputs and outputs without prior assumptions (Naser and Alavi 2021). Regression can be performed using different ML algorithms, such as neural networks and decision trees (Salehi and Burgueño 2018).
ANN has been extensively used in engineering for capturing the relationship between input and output variables involved in a system (e.g., Naderpour et al. 2018; Naser et al. 2012). ANN can be efficiently applied for predicting the response of beam–column connections that can be used efficiently for the simulation of self-centering building structures. The first application of ANN for predicting the response of steel connections dates back to the ‘90s when Abdalla and Stavroulakis (1995); Stavroulakis et al. (1997) estimated the mechanical behavior of semirigid connections and proposed a moment-rotation constitutive law using neural networks. Anderson et al. (1997) applied ANN for predicting the bilinear moment-rotation response for minor-axis connections. Yun et al. (2008) proposed a design-variable-based inelastic hysteretic model for beam-to-column connections using neural networks. Their results demonstrated that the proposed module could reproduce the experimental data with reasonable accuracy for extended endplate connections and top-and-seat angles with double-web angle connections. In a study by Shah et al. (2018), 32 boltless connections were tested to propose predictive models for the moment-rotation response of the connections. For developing predictive models, they used three methods, including linear genetic programming, ANN, and an adaptive neurofuzzy inference system. Horton et al. (2021) predicted the cyclic response of reduced beam section (RBS) connections using neural networks. Recently, Kim et al. (2021) used ML techniques, including support vector regression (SVR) and deep neural networks, to predict the strengths of steel circular hollow section X-joints. Their results showed that deep neural networks provided more accurate results compared with SVR.
Despite the fast-growing interest in using AI and ML techniques, their application for predicting the self-centering response of SMA-based steel connections and frames has not been widely explored. This paper aims to develop predictive models for the response of SMA-based extended endplate connections using ANN. The proposed models are then implemented to develop a MATLAB tool for response prediction of SMA-based connections. The study is then expanded to optimize SMA-based connections to improve the connection response characteristics while minimizing the amount of steel and SMA materials. Next, the developed predictive tool is used to create a computationally efficient and accurate phenomenological model for the SMA connections in OpenSees version 3.3.0. Finally, the accuracy of the developed predictive modeling and optimization results is confirmed.

Scope and Methodology

In this study, a design database from FE simulation and experimental tests of SMA-based extended endplate connections (Fang et al. 2013) was used to propose predictive models for the connections’ moment-rotation response parameters, including θB, θC, θE, MB, MC, ME, and β as shown in Fig. 2. This figure shows a schematic view of the idealized response curve [i.e., self-centering response and backbone curve consistent with the generalized backbone curve in ASCE 41-17 (ASCE 2017)]. These response parameters were selected to idealize the backbone and self-centering response of endplate connections with SMA bolts on the basis of experimental (e.g., Fang et al. 2013) and FE analysis results (e.g., Mohammadi Nia and Moradi 2020). To create a response curve for the connections, SMA materials were assumed to fracture upon reaching strain εfr (Mohammadi Nia and Moradi 2021; Sabouri Ghannad et al. 2021):
εfr=εL+σMfESMA
(1)
where εL = maximum transformation strain of SMA bolts; σMf = martensite finish stress; and ESMA = modulus of elasticity of SMA material. SMA bolts fracture is more likely to occur at a strain level around εfr (Wang et al. 2017).
Fig. 2. Idealized response curve for extended endplate connections with SMA bolts.
In the proposed backbone curve, as shown in Fig. 2, from point A to B, the connection has an elastic behavior. From point B onward, the SMA material enters a forward transformation phase. At point C, the outmost SMA bolts reach their fracture strain, which in turn results in a drastic reduction in the moment capacity of the connection. Following point C, the loading continues to achieve the fracture strain in the second row of the SMA bolts. Additionally, a return path is included using β to characterize the self-centering response.
Ten factors have been identified as influential factors in the cyclic response of SMA-based connections (Mohammadi Nia and Moradi 2020). These influential factors are listed in Table 1. The factor ranges were selected while ensuring that any developed connection using these factor ranges meets the seismic requirements of the American Institute of Steel Construction [AISC 341-16 (AISC 2016b); AISC 358-16 (AISC 2016a)] for extended endplate connections. The endplate thickness has a constant value of 40 mm to ensure the thick-plate behavior by the endplate. The slenderness ratio of the beam and column flanges and webs were considered based on the limits for highly ductile sections as per AISC 341-16 (AISC 2016b). Additionally, the requirements in AISC 358-16 (AISC 2016a) for extended endplate connections were checked, including the restrictions or requirements for beam and column sizes, beam clear span-to-depth ratio, lateral bracing for beams and columns, and the strong-column-weak-beam requirement. Moreover, the column dimensions were set to have strong-column-weak-beam behavior with no need for doubler and continuity plates as per AISC 341-16 (AISC 2016b). Column web and flange thicknesses are 30 mm and 35 mm, respectively. Based on the results of a statistical sensitivity analysis (Mohammadi Nia and Moradi 2020), column-related factors do not significantly affect the response characteristics of the connection. This is also expected because the column and particularly the panel zone is designed to remain elastic ensuring that large deformations are confined to the SMA bolts. Therefore, considering the same column web and flange thicknesses in all the connections should not impact the conclusions from the study. With these significant factors, a data set of 72 factor combinations generated using the design of experiments (Mohammadi Nia and Moradi 2021) was considered. For each factor combination, two analyses were performed in ANSYS mechanical APDL (ANSYS 2020). In the first analysis, the connection was loaded up to 0.08 rad rotation to capture the rotation at which the outermost bolts reach their fracture strain as defined in Eq. (1). A second analysis was next performed on the same FE model of the connection but without the outmost (i.e., first-row) SMA bolts. In fact, the FE model represents a connection with its outmost SMA bolts fractured (thus removed from the analysis). The connection model was loaded up to 0.10 rad rotation to capture the rotation at which second-row SMA bolts were fractured. From the second analysis, the residual strength of the beam–column connection (indicated by line DE in Fig. 2) was obtained. A backbone curve similar to the one in Fig. 2 was obtained for each factor combination by assembling the backbone curves achieved from FE simulations.
Table 1. Factors and ranges considered
FactorSymbolLow levelHigh levelUnit
Martensite start stressσMs280380MPa
Martensite finish stressσMf410590MPa
Austenite start stressσAs170250MPa
Austenite finish stressσAf70138MPa
Maximum transformation strainεL0.070.13
Bolt pretension strainεpt0.0050.015
Bolt lengthLbolt300350mm
Bolt diameterDbolt1025mm
Beam depthhbeam150610mm
Beam lengthLb1,5004,500mm
In addition to the data set developed using the design of experiments and FE simulations, results from seven experimental tests performed by Fang et al. (2013) were considered. For the specimens experiencing early fracture in the experimental tests (due to their low net threaded-to-shank diameter ratio of 1), θC was not quantified.
Next, ANN was used to develop a predictive model for the response parameters for SMA-based extended endplate connections. The inputs to the neural networks were those ten design parameters listed in Table 1. To characterize the backbone and self-centering response, a network was trained for each response parameter, including θB, θC, θE, MB, MC, ME, and β. For each response parameter, separate ANNs were developed, resulting in higher prediction accuracy in contrast to training a single ANN for all the response parameters. The trained networks were then assembled to have a graphical interface for predicting the response.
Additionally, to study the simultaneous response improvement and cost-effectiveness of the SMA-based connections, multiobjective optimizations were performed. A multiobjective evolutionary algorithm based on Pareto-dominance was adopted in which the selected objectives were optimized simultaneously. A desirability approach was adopted to rank the optimal solutions in a Pareto front set and thereby to find and propose the solution with the highest rank as the optimal factor setting. The desirability approach was also used to propose optimal ranges for factors of interest, including beam depth and SMA bolt length and diameter. Two optimization problems were defined to improve the response of SMA-based connections and reduce the amount of SMA and steel materials, thus resulting in cost-effective connections with improved performance.
Using the proposed backbone curve, a phenomenological model was developed and validated for two experimentally tested SMA-based connections. The efficiency and accuracy of the developed MATLAB tool and proposed phenomenological models were confirmed using two beam-to-column connections. For each connection, the backbone curve was developed using the MATLAB tool, and then, fiber and detailed FE models were developed in OpenSees and ANSYS, respectively. The hysteretic responses obtained from ANSYS and OpenSees, along with predicted backbone curves, were compared. Finally, system-level modeling was performed. Pushover and nonlinear response history analyses were performed to confirm the optimization study results and to assess the connections’ behavior in steel frames.

FE Modeling and Validation

Three-dimensional FE models were developed using ANSYS mechanical APDL (ANSYS 2020), as shown in Fig. 3(a). SOLID185 elements were used to model the beam, column, SMA bolts, and stiffeners. SOLID185 has eight nodes with three degrees of freedom at each node: translations in the nodal x, y, and z directions. Finer meshes were used in the connection interface, in which inelastic deformations may occur, whereas for the rest of the components, coarse meshes were assigned to have efficient and yet accurate FE models. CONTA174 and TRAGE170 were used to define contact surfaces between the endplate and column flange, as well as the probable contact between SMA bolts and holes within the endplate and column flange. In defining contacts, the penalty function was used as the contact algorithm, while the stiffness of the contact was set to be updated in each iteration according to the mean stress of the underlying elements. Superelastic behavior was adopted for SMA materials, while trilinear behavior was used to model steel materials (Mohammadi Nia and Moradi 2020). Large deformation and nonlinear materials were activated in the developed models to account for geometric and material nonlinearities, respectively. In the FE models, geometric imperfections and the effect of residual stresses were not included. In order to avoid convergence difficulties, the number of substeps for the load step, in which the convergence issue could occur, was increased to apply the load slowly. Furthermore, an unsymmetric Newtown–Raphson algorithm was used to avoid convergence difficulties (Moradi and Alam 2015b). ANSYS default convergence tolerances of 0.5% and 5%, respectively, were adopted for force and displacement. The moment-rotation response from the FE models was compared with experimental results [Fig. 3(b)]. A good agreement between the FE results and experimental results was observed. Further details of the FE modeling and validation studies are available elsewhere (Mohammadi Nia and Moradi 2020, 2021).
Fig. 3. FE modeling and verification with specimen D10-240d: (a) the developed FE model; and (b) comparison between moment-rotation response. (Data from Fang et al. 2013.)

Design Database Generation

From the design of the experiment method, an I-optimal design was used to efficiently generate the design database for the predictive modeling development. The generated database constitutes sampling points or factor combinations with the factors and associated ranges in Table 1. In optimal designs, a numerical criterion, i.e., the variance or any statistical properties of the design, is optimized. An I-criterion is one of the most well-known designs that is used when the experimental goal is to make precise predictions of the response rather than to obtain precise estimates of the model parameters (Smucker et al. 2018). An I-optimal design is more appropriate for developing predictive models based on computer simulations. In this design, a set of runs, i.e., factor combinations, are selected that minimize the integral of the prediction variance across the factor space (Montgomery 2006).

Artificial Neural Network Predictive Models

Inspired by the human nervous system, ANN is a computational model that contains multiple layers of artificial neurons connected with coefficients (weights). ANN can model complex problems with many parameters when being trained by proper training data (or exemplars). ANN consists of interconnected neurons, which are processing elements with similar characteristics such as input, synaptic strength, activation output, and bias. The processing units, including input, hidden, and output layers, carry the weights of the network. The training process of a network is associated with adjusting the weights to optimize the desired loss function, which is usually defined as the prediction error.

ANN for Predicting Response Parameters

ANNs were used to determine the nonlinear relationship between the input and output parameters of the response curve for SMA-based extended endplate connections. To that aim, a single layer perceptron (i.e., a feed-forward network with a single hidden layer) was used, as shown in Fig. 4. The backpropagation algorithm was used to train the network where the output errors were propagated back by means of the same connections used in the feed-forward mechanism. In this method, the gradient of the error with respect to the weights is calculated, and then the weights are updated using the popular gradient descent algorithm. In the proposed neural networks, the neurons were placed in three separate layers, including the input layer, hidden layer, and output layer. The input layer has ten neurons, reflecting 10 design parameters as listed in Table 1. Neurons in the input layer pass the scaled input data to the hidden layer using weights. The hidden layer has a different number of neurons for each response variable, obtained using trial and error to result in high accuracy. The number of hidden layers was changed from 1 to 3, while the number of neurons in the hidden layers was changed from 1 to 20—in searching for the best (or optimal) ANN with the minimum observed mean squared error. Using a single hidden layer for all the trained networks outperformed models with 2 or 3 hidden layers. Table 2 lists the number of neurons used in the trained ANNs. The training was repeated with different numbers of hidden layers and neurons until the best ANNs were found. The optimal ANNs were then used for developing the predictive tool.
Fig. 4. Schematic representation of a single hidden layer feed-forward neural network.
Table 2. The details of the ANN models
Response parameterNumber of neurons in the hidden layerRR2
AllTrainTestAllTrainTest
θB50.960.960.950.920.930.91
θC70.980.990.950.960.980.91
θE70.990.990.980.970.990.96
MB51.001.000.990.990.990.99
MC70.991.000.990.991.000.99
ME70.991.000.990.991.000.99
β30.980.990.980.970.970.97
The output layer comprised a single neuron that represents the backbone curve parameter. In this study, the hyperbolic tangent sigmoid transfer (tansig), i.e., g(x)={2/[1+e(2x)]}1, which generates an output between 1 and 1, was used as an activation function for the hidden layer, while the pure linear function (purelin), which generates outputs between and +, was used as the activation function in the output layer. The goal of using nonlinear transform functions like the tangent sigmoid transfer is to provide the network with the capability of learning the nonlinear behavior between the input and output data. Since the networks in the study were shallow networks, we used the tangent sigmoid function (Szandała 2021). As shown in Fig. 4, each neuron in a layer is connected to the neurons in the next layer, whereas there is no interconnection between neurons.
The design matrix developed using the I-optimal design, along with the experimental data from Fang et al. (2013), was used as the input matrix for neural networks. The neural network tool in MATLAB was utilized to develop the multilayer perceptron architecture of feed-forward ANN. The developed design matrix using the I-optimal design and the experimental data contains 79 factor combinations from which 70% of the data were used to train the networks, 15% of the data was used for validation, and the last 15% of the entire database was used for testing the networks. The performance of the networks was evaluated using the coefficient of correlation (R) and coefficient of determination (R2). Note that R, which estimates the relationship between model output and actual values, ranges from 1 to 1. For R and R2, close-to-one values are desirable. The coefficient of correlation (R) measures the strength of the relationship between input and output data. R was used to find patterns and relationships in the data. Further, the coefficient of determination (R2) shows how well the regression model fits the observed data. R2 was used to evaluate predictions and see how much of the variation in the FE and experimental data is explained by ANNs. It should be mentioned that mean squared error (MSE), measuring the average of the squares of the errors (Naser and Alavi 2021), was used implicitly in the training process of the ANNs. MSE was used as the performance metric for the validation data. Minimum MSE for the validation data was considered as a measure to select the weights in the ANNs and as an indicator for the best performance of the ANNs. The input data for all neural networks were normalized to lie within a range of 0 to 1. The reason for normalizing the data was to prevent the saturation of the tansig activation function, which could stop the network from further learning (Fausett 2006). Eq. (2) is used to normalize the inputs of neural networks
φm=xxminxmaxxmin
(2)
where φm = normalized data; and x = original data.

ANN Results

For each response parameter, an optimized feed-forward neural network was trained using the Levenberg–Marquardt algorithm. This training process automatically stops as soon as the generalization stops improving, and this is signaled by the increase of MSE in the validation samples. The number of epochs is kept minimum during the development of the optimal topology of the networks to avoid overfitting where the network perfectly fits the training data, but its performance for the testing set is not satisfying.
Based on the FE analysis results and experimental data, seven neural networks were trained (one for each backbone curve parameter). Table 2 lists the number of neurons that were used in the hidden layer for each neural network associated with different backbone parameters. As listed, the coefficient of correlation, R, between actual and predicted values for the trained neural networks is quite high, ranging from 0.96 to 1.0. Moreover, the coefficient of determination, R2, is calculated for the trained networks. R2 measures the proportion of the variance in the response variable that is predictable from the trained ANN. R2 ranges from 0.92 to 0.99.
Fig. 5 presents the predicted versus actual responses for the training, testing, and validation sets for different response curve parameters. As shown, the data are clustered along the diagonal line, indicating the accuracy of the trained neural networks.
Fig. 5. Target data versus predicted response in the training, validation, and test sets for (a) θB; (b) MB; (c) θC; (d) MC; (e) θE; (f) ME; and (g) β.
Following the framework by Naser et al. (2021), a benchmarking study is presented to apply six commonly used supervised learning algorithms with default settings to the database for SMA connections. The algorithms include decision trees (DT), random forest (RF), extreme gradient boosted trees (ExGBT), light gradient boosted trees (LGBT), TensorFlow deep learning (TFDL), and Keras deep residual neural network (KDP). For consistency with the trained ANNs in the paper, one model was trained for each response parameter. R2 and root MSE (RMSE) were used as the performance metrics as listed in Table 3. Note that the trained algorithms are not finetuned (Naser et al. 2021). Comparing the performance of the algorithms (while excluding the trained ANNs) shows that no single algorithm outperforms all the examinations. However, by considering the overall performance of the algorithms, ExGBT and KDP are the first and second best-ranked algorithms, while TFDL and DT have the poorest performance. In general, the trained ANNs in this study outperform other algorithms; nonetheless, it should be noted that the algorithms in Table 3 are not finetuned and this comparison is not fair. Finetuning the algorithms could result in more accurate predictions. Further research is warranted to find the best algorithm for each response parameter.
Table 3. Performance of the trained ML algorithms using the developed database
ParametermetricDTRFExGBTLGBTTFDLKDPANN
TrainTestTrainTestTrainTestTrainTestTrainTestTrainTestTrainTest
θBR21.000.790.890.781.000.880.920.860.860.680.790.520.930.91
RMSE0.000.00100.00080.00140.00010.00100.00070.00110.00100.00160.00120.00150.00070.0010
θCR21.000.550.870.741.000.820.940.910.610.420.930.890.980.91
RMSE0.00060.00770.00410.00730.00070.00450.00310.00300.00870.00900.00300.00400.00200.0040
θER21.000.760.910.911.000.860.950.930.640.280.930.920.990.96
RMSE0.000.0100.0050.0050.000.0060.0040.0040.0080.0180.0050.0040.0010.004
MBR21.000.980.970.921.000.990.980.970.920.901.000.990.990.99
RMSE033.435.560.51.223.332.342.158.470.312.417.214.917.0
MCR21.000.940.980.961.000.990.980.980.870.711.000.991.000.99
RMSE0.3772.937.756.61.026.843.141.0102.5142.38.317.26.326.4
MER21.000.950.970.941.000.970.970.940.970.911.001.001.000.99
RMSE136.025.132.20.729.023.645.225.655.06.09.39.09.1
βR21.000.900.910.841.000.900.920.910.730.620.910.890.970.97
RMSE0.000.030.030.050.00.030.030.030.060.060.400.300.020.02

MATLAB Graphical User Interface

The trained neural networks were used to develop a MATLAB graphical user interface (GUI) (Mohammadi Nia and Moradi 2022) for the response prediction of SMA-based extended endplate connections. When the goal is to only determine the moment-rotation response, the developed tool eliminates the need for a detailed FE modeling of self-centering connections and frames. The tool is also useful for the efficient analysis of frames with SMA connections. As shown in Fig. 6, 10 significant parameters are required to be entered into the developed tool. Note that the developed predictive tool is applicable only for the factor ranges listed in Table 1. In developing the predictive tool, the beam flange (tbf=19  mm), beam web (tbw=15  mm), column flange (tcf=35  mm), and column web (tcw=30  mm) thicknesses were kept constant. These factors do not significantly influence the connection response based on statistical sensitivity analysis results (Mohammadi Nia and Moradi 2020)—given that the slenderness ratios for beam and column flanges and webs meet the requirements of AISC-341-16 (AISC 2016b) for highly ductile sections and the requirements of AISC-358-16 (AISC 2016a) for extended endplate connections. As a result, in properly designed SMA connections, the beam and column remain elastic, and the connection response will not be affected by using different flange or web thicknesses for beams and columns. An unloading path is also included to facilitate implementing the proposed response curve in other software programs, such as OpenSees, which is discussed subsequently in this paper.
Fig. 6. Developed MATLAB tool for predicting the response of SMA-based endplate connections.

Multiobjective Response Optimization

The proposed ANNs were used to perform response optimization studies. Optimal factor settings were determined to improve the connection response characteristics, such as the initial stiffness and the moment and rotational capacities with minimized steel and SMA material usage. The input space of the neural networks was optimized using a genetic algorithm. A general multiobjective optimization problem can be formulated as follows (Deb 2011):
Minimize/Maximize fm(x)
Subjectto  gj(x)0,hk(x)=0,xi(L)xixi(U)
(3)
m=1,2,,M;j=1,2,,J;k=1,2,,K;i=1,2,,I
where x = vector of decision variables containing the input factors, that is, (σMsσMfhbeamLb)T. The functions gj(x) and hk(x) are inequality and equality constraint functions, respectively; M = number of optimization objectives; and I = number of decision variables. Further, J and K are the numbers of constraints. The constraints for xi are variable bounds based on the factor ranges in Table 1.
In multiobjective optimization problems, there is no unique solution; rather, there is a set of optimal trade-off solutions called the Pareto front (Ngatchou et al. 2005). A solution belongs to a Pareto set, given that there exists no other solution that can improve at least one of the objectives without deteriorating any other objectives. The optimal solutions have no mutually dominated relationship. In addition, any solution outside the Pareto front is always dominated by a solution in the Pareto front. Fig. 7 illustrates a schematic view of the Pareto front for a two-objective optimization problem with different objectives where the Pareto frontiers are highlighted in red.
Fig. 7. Pareto front for a two-objective optimization problem when: (a) maximizing both objectives; and (b) maximizing one objective and minimizing the other.
In this study, MATLAB solver gamultiobj, which is a controlled elitist genetic algorithm, is adopted. In a controlled elitist genetic algorithm, individuals that can help to increase the diversity of the population (a subset of solutions) are preferred even with a lower fitness value, in which the fitness value shows the accuracy level of the solution.

Optimization Objectives

Two multiobjective optimization problems were considered. The first optimization problem was defined to enhance the response characteristics of SMA-based steel connections while reducing the amount of SMA and steel materials. The second optimization, however, only seeks to improve the response of the connection regardless of the amount of material.
In the first optimization problem, the amount of SMA and steel materials were minimized in addition to maximizing the initial stiffness and the deformation and moment capacities (θC and MC) of the connection. The optimization was set to increase the connection’s initial stiffness knowing that reducing the flexibility of moment-resisting frames (MRFs) is favorable during earthquakes. Given the inherent low lateral stiffness of MRFs, buildings experience higher lateral drifts. As a result, reducing the flexibility of MRFs is favorable during earthquakes. Using SMAs, which possess a lower modulus of elasticity compared to steel, exacerbates the low lateral stiffness of the system. Therefore, increasing the lateral stiffness of MRFs, particularly with SMA-based connections, is favorable and needed.
Moreover, the amount of materials used in a beam–column connection contributes to the total cost of the connection. Hence, the first optimization problem was defined to minimize the amount of steel (i.e., the beam depth) and SMA (i.e., bolt length and diameter).
The second optimization problem seeks solutions that improve the response characteristics of the SMA-based endplate connection by maximizing the rotational capacity (θC) and the moment capacity (MC). Increasing the deformation capacity of SMA-based beam-column connections is important for providing adequate ductility. Past research has shown that early SMA bolt fracture results in lower deformation capacity of the SMA-based steel connections (Fang et al. 2013).

Optimal Point Selection and Proposed Optimal Ranges

Desirability functions were used to select the most optimal solution and to prioritize the points in a Pareto optimal set. In this ranking method, each objective is normalized using the following equations for maximization and minimization problems, respectively:
Inmaximizationproblems:d={0y<ymin(yyminymaxymin)yminyymax1y>ymax
(4)
Inminimizationproblems:d={1y<ymin(ymaxyymaxymin)yminyymax0y>ymax
(5)
where d = normalized objective; y = obtained objective in the specific Pareto point; and ymax and ymin = maximum and minimum responses, respectively. Close-to-one normalized objectives (d) have priority. The following equation is used to calculate the desirability for each optimal solution in the Pareto optimal set:
D=(d1×d2××dn)1n
(6)
where D = desirability score; and n = number of objectives. The desirability score for each point was calculated using Eqs. (4)(6). The points with the highest desirability scores were obtained, which were then used to define optimal ranges for the factors in the optimization problem. In finding optimal factor ranges, the factor of interest was changed from its minimum to maximum to calculate the disability score at different levels of the factor. A range in which the desirability score was at least 85% of the maximum observed desirability was selected as the optimal range.

Optimization Results

Results of the First Optimization Problem

The first optimization problem aims to reduce steel and SMA material usage while enhancing the structural response characteristics of the SMA-based connection by increasing the moment and rotational capacities and initial stiffness [Ki=(MB/θB)]. In frames with stiffer connections, the lateral displacement under seismic loading is reduced. Additionally, to account for minimizing the amount of steel and SMA materials, the optimization was set to minimize the SMA bolt diameter, SMA bolt length, and beam depth.
Fig. 8 presents the Pareto frontier in the first optimization problem. These plots show the variation of the Pareto frontiers with respect to the bolt diameter, bolt length, and beam depth. As shown, higher bolt diameters lead to higher initial stiffness and moment capacity, whereas the deformation capacity (represented by θC) is not changed with the variation in the bolt diameter. In Fig. 8, it is observed that the bolt length mainly influences the deformation capacity of the connection. The longer the SMA bolt, the higher the rotational capacity. Additionally, connections with deeper beams provide higher Ki and MC.
Fig. 8. Variation of the Pareto optimal set in the first optimization problem with respect to (a) bolt diameter; (b) bolt length; and (c) beam depth.
The desirability approach was utilized to select the most optimal points from the Pareto optimal set. Table 4 lists three factor settings with the highest desirability score in the first optimization problem. Solution 1 in this table was used to determine optimal ranges for the factors that were considered in the optimization objectives (i.e., the SMA bolt diameter and length and the beam depth). To that aim, the factor of interest was changed between its minimum and maximum, while other factors were unchanged in the optimal factor combination. The variation of the desirability function with respect to the bolt diameter, bolt length, and beam depth is illustrated in Fig. 9. Since the optimization goal was to minimize these factors, the desirability function was drastically reduced to zero once these factors were at their maximum value. The optimal ranges are listed in Table 5. These optimal ranges result in the top 15% desirability (i.e., higher than 0.85×0.47=0.4).
Table 4. Points with the highest desirability score in the first optimization problem
Solution numberσMs (MPa)σMf (MPa)σAs (MPa)σAf (MPa)εLεptLbolt (mm)Dbolt (mm)hbeam (mm)Lbeam (mm)Desirability
13295092081180.1170.00919522.32571,9890.47
23415782071130.1080.01121818.63082,1010.45
33555672211110.1060.00921720.63871,9040.45
Fig. 9. Variation of the desirability function with respect to (a) bolt diameter; (b) bolt length; and (c) beam depth.
Table 5. Optimal ranges found in the first optimization
FactorMinimum valueMaximum valueUnit
Dbolt1624mm
Lbolt190265mm
hbeam150400mm

Results of the Second Optimization Problem

In the second optimization problem, the design space was searched for maximizing the rotational and moment capacities of the SMA-based steel connections regardless of the amount of SMA and steel materials. Fig. 10(a) shows a set of optimal solutions from the second optimization problem. The trade-off that exists between θC and MC is clearly shown in Fig. 10(a). Based on the Pareto optimal solution set, a 125% increase in θC from 0.04 rad to 0.09 rad reduces the moment capacity by 34% from 872  kN·m to 578  kN·m. To illustrate the variation between MC and θC, three points are denoted with A, B, and C in the Pareto front diagram. These three points correspond to the beam–column connections that provide the highest MC, a trade-off between MC and θC, and the highest θC, respectively. The moment-rotation curves for the connections corresponding to points A, B, and C are shown in Fig. 10(b). Generally, Pareto optimal frontiers can be used to select the point (corresponding to the beam–column connection) suiting any specific need, such as high moment capacity or deformation capacity.
Fig. 10. Second optimization: (a) Pareto optimal frontiers; and (b) moment-rotation curves for the beam–column connections represented by points A, B, and C.
As per AISC 341-16 (AISC 2016b), connections in special moment frames are required to accommodate a story drift angle of at least 0.04 rad. As shown in Fig. 10(a), the rotational capacity of the optimal solutions is more than 0.04 rad. Therefore, the optimal connections meet the requirements of AISC 341-16 in terms of the rotational capacity for special moment frames.
By exploring the Pareto set in the second optimization problem, it is found that the SMA bolt diameter for the entire Pareto set has its maximum value of 25 mm. By increasing the SMA bolt diameter, the axial force capacity of SMA bolts is increased, resulting in connections with higher moment capacity. Additionally, it is noticed that 96% of the Pareto optimal solutions have the maximum transformation strain of 0.13. The fracture of SMA bolts is correlated with the maximum transformation strain of SMA; therefore, SMAs with higher values of maximum transformation strain are desirable to avoid early bolt fracture and consequently provide the connection with higher deformation capacity.
The desirability approach was applied to select the most optimal solution for which the maximum desirability score was obtained. As listed in Table 6, the bolt diameter and maximum transformation strain are at their high levels, whereas lower pretension strains (close to 0.005, the low level for εpt) are found as the optimal solution. The latter can be attributed to the fact that increasing the prestrain level in SMA bolts increases the likelihood of early bolt fracture, reducing the deformation capacity and thus the desirability. In Table 6, the desirable point has a σMs close to its minimum value of 280 MPa, whereas σMf is close to its maximum value of 590 MPa. Therefore, the larger difference between σMs and σMf is found to be desirable to reach higher strength and deformation capacities. This finding is consistent with results from past research indicating that the higher slope of the response curve (governed by the SMA forward transformation; line BC in Fig. 2) is favorable and cost-effective (Qiu and Zhu 2017).
Table 6. Optimal factor setting from the results of the second optimization study
FactorOptimal value
σMs (MPa)289
σMf (MPa)578
σAs (MPa)223
σAf (MPa)115
εL0.129
εpt0.007
Lbolt (mm)305
Dbolt (mm)25
hbeam (mm)582
Lbeam (mm)3,824

Note: desirability = 0.723.

The most optimal factor setting, as listed in Table 6, was used to obtain the optimal factor ranges. The variation of the desirability function with respect to each design parameter was used to select optimal ranges resulting in the top 15% desirability. These optimal ranges in the second optimization problem are listed in Table 7. As shown, to obtain the maximum moment and rotation capacities, the SMA bolt diameter and maximum transformation strain should be at high levels, while the bolt pretension strain should be at a low level.
Table 7. Optimal factor ranges obtained in the second optimization problem (maximizing moment and rotational capacities)
FactorMinimum valueMaximum valueUnit
σMs280380MPa
σMf576590MPa
σAs194250MPa
σAf70117MPa
εL0.1280.13
εpt0.0050.007
Lbolt300350mm
Dbolt2525mm
hbeam560610mm
Lb3,5404,500mm

Developing FE Models using the Proposed Predictive Tool

This section presents the development and verification of a phenomenological model for SMA-based extended endplate connections. The proposed OpenSees model, along with the MATLAB predictive tool, was then used to develop high-fidelity FE models for steel frames equipped with SMA-based connections.
OpenSees (2013) was used to develop a simple yet efficient numerical model for SMA-based extended endplate connections. The connection was modeled using rotational springs with Self-Centering, Pinching4, and Steel01 material models acting in parallel. Fig. 11 shows the proposed phenomenological model. The accuracy and effectiveness of the developed model were verified by modeling the experimentally tested connection specimens SMA-D10-240d and SMA-D10-290 (Fang et al. 2013). The material properties for the Self-Centering, Pinching4, and Steel01 models were selected using experimental backbone curves (Fang et al. 2013). An iterative process was used to adjust the percentage contribution of the material models. Through this iteration, a linear combination of Self-Centering, Pinching4, and Steel01 was found and assigned with the coefficients of 0.90, 0.05, and 0.05, respectively. In the proposed phenomenological model, the beam and column were modeled using elastic beam–column elements. As shown in Fig. 12, there is a good agreement between the OpenSees and experimental results.
Fig. 11. Schematic view of the phenomenological model in OpenSees.
Fig. 12. Comparison between the results of the proposed OpenSees model and experimental data for specimens: (a) SMA-D10-240d; and (b) SMA-D10-290. (Data from Fang et al. 2013.)
After verifying the OpenSees modeling, we demonstrate the efficiency of the developed MATLAB version R2020b tool and confirm its prediction accuracy. For this purpose, the MATLAB tool was first used to predict the backbone response for two beam-column connections, including a connection with randomly selected details and the most optimal connection found in the second optimization study. The predicted response was next compared with the numerical results from the fiber-based and detailed solid models using OpenSees and ANSYS, respectively. For each beam-column connection, the backbone curve was developed using the predictive MATLAB tool. The backbone curve parameters from the developed MATLAB tool were then introduced in OpenSees to create fiber-based models. Fig. 13 shows the hysteretic response of the developed ANSYS and OpenSees models along with the predicted backbone curve. The response predictions using the MATLAB tool and the OpenSees modeling agree well with the cyclic response from the detailed ANSYS models.
Fig. 13. Verification results for: (a) a randomly selected connection; and (b) the optimal connection in the second optimization.
The results of the verification study in Fig. 13 also confirm the insignificant influence of the panel zone. The response from the OpenSees model matches that from ANSYS, which explicitly captures any deformation and inelasticity in the panel zone. In fact, SMA-based connections are designed for maximum recentering capability. As a result, the beam and column (including the panel zone) remain elastic to confine large deformations in the SMA bolts and thus prevent any damage to other components of the connection. Moreover, based on the results of a statistical sensitivity analysis (Mohammadi Nia and Moradi 2020), the panel zone-related factors are not influential on the hysteretic response of the connections that are designed to confine deformations in the SMA bolts. Therefore, given the design with strong panel zones, their effect on the response is insignificant.
The described OpenSees modeling, along with the response prediction from the MATLAB tool, greatly facilitates the efficient analysis of self-centering connections and frame structures. This efficiency becomes notable when comparing the analysis runtime of the ANSYS and OpenSees models. For a typical beam–column connection, an ANSYS model with 80,631 elements and 92,705 nodes requires a total runtime of 6 h and 48 min using a 64-bit operating system in Windows 10 with a six-core Intel Xeon W-2135 CPU at 3.70 GHz and with 16 GB of RAM. In contrast, the OpenSees analysis for the same connection takes only 3 min to complete while providing the same level of response prediction accuracy as in the ANSYS model. This demonstrates the efficiency of using the predictive models to significantly reduce the runtime while maintaining accuracy. However, it should be noted that the phenomenological model only predicts the moment-rotation response of the connection. Detailed continuum FE modeling, in contrast, provides additional information that is not available from the simplified phenomenological model.
Note that the use of the design of experiments method for generating the factor combinations can make the study more efficient. The verification results in this section indicate that 79 samples are sufficient to train ANNs with high prediction accuracy.

System-Level Modeling

This section illustrates the use of the developed predictive tool for modeling steel moment frames with SMA-based connections. Further, the optimization results were confirmed by evaluating the system-level response of frames equipped with optimized beam-column connections.

Structural Modeling of Moment Frames

A 3-story single-bay MRF was used for the illustration. The bay width is 4 m, and the first and typical story heights are 4.6 m and 4 m, respectively. A seismic mass of 94  kN-sec2/m, 103  kN-sec2/m, and 104  kN-sec2/m was considered for the first, second, and third floors, respectively. Two-dimensional nonlinear structural models were developed using OpenSees (2013), as shown in Fig. 14. The beams and columns were modeled using force-based elements. The beam-to-column connections were developed as discussed previously. A leaning column was used to support the building weight and generate P-Δ effects as the result of lateral deformations. Truss elements were used to connect the leaning column to the MRF to avoid any contribution to the lateral load resistance.
Fig. 14. Numerical model of three-story frame.
As listed in Table 8, four-moment frames were considered with randomly selected factors for their SMA connections. Unoptimized frames 1 and 2 have SMA connections that were set with randomly selected factors from the original factor ranges. However, optimized frames 3 and 4 have SMA connections with factor settings that were randomly selected from the optimal ranges for maximizing the moment and rotation capacities (listed in Table 7). Identical beams (W27×102) and columns (W24×131) were considered for all the frames to provide the basis for comparing the effects of the SMA connections. The beam and column sizes were selected based on the capacity design to ensure they remained elastic.
Table 8. Randomly selected factor settings for the beam-to-column connections in each frame
FrameσMs (MPa)σMf (MPa)σAs (MPa)σAf (MPa)εLεptLbolt (mm)Dbolt (mm)hbeam (mm)Lbeam (mm)
12935292191100.1000.005329125801,878
2356470189850.0700.011311145172,129
3311586214790.1300.007327255783,834
42835882431070.1280.007330256094,173

Response Evaluation: Pushover and Response History Analysis

To compare the influence of the SMA connections on the frame lateral load and deformation capacities of the MRFs, we performed pushover analyses on the frames. A roof drift target of 10% was considered for the pushover analysis. Fig. 15(a) presents the pushover curves for the frames. As shown, the optimized frames 3 and 4 possess greater lateral load and drift capacities, as expected owing to their optimal connections. The SMA bolt fracture in the optimized frames occurs at a drift level of 6.5%, whereas the rotational capacity of the unoptimized frames can be as low as 3%. This confirms that the proposed optimal ranges are effective in maximizing the lateral load resistance and drift capacity of self-centering steel frames.
Fig. 15. System-level assessment of optimized and unoptimized frames: (a) pushover curves; and (b) residual story drifts under different earthquake records.
Nonlinear response history analyses were also performed on the frames using a suite of 20 ground motion records (LA21–LA40) developed by Somerville (1997) for Los Angeles. These ground motion records represent the maximum considered earthquake (MCE) intensity level. Note that the choice of the earthquake records was not critical in this study because our objective was to illustrate the modeling approach and compare the seismic performance of the SMA-based frames with different SMA beam–column connections.
The maximum residual story drift was evaluated for each frame under the earthquake records with additional 25 s of free vibration. The free vibration was considered at the end of the earthquake loading to extend the time integration in the analysis to accurately capture residual rotations. Fig. 15(b) shows the record-to-record and mean residual deformations of the optimized and unoptimized frames. With an average residual story drift of 0.04%, the optimized frames exhibit excellent self-centering.
Fig. 16 compares the rotation history of the first-floor connections of Frames 1 and 3, which experience a maximum rotation of 0.028 rad and 0.062 rad, respectively, under earthquake record LA25. In contrast to the full recentering of the optimized connection, the residual rotation of the unoptimized connection in Frame 1 is evident. This residual deformation is due to the SMA bolt fracture in the connection because the rotation exceeds its capacity (0.028 rad). Therefore, using the determined optimal ranges can improve the self-centering response of SMA connections by delaying the SMA bolt fracture. Similar results were found when using different earthquake records. As expected, the plastic deformations of the beams and columns were almost zero. The average residual rotation for the optimized and unoptimized frames were 0.04% and 0.46%, respectively, as shown in Fig. 15(b).
Fig. 16. Rotation history of the first-floor connections under earthquake record LA25: (a) unoptimized connection in Frame 1; and (b) optimized connection in Frame 3.

Summary and Conclusions

In this study, ANNs were used to develop predictive models for self-centering response and backbone curve parameters of extended endplate connections with SMA bolts. The predictive models were developed based on an input database comprising results from 72 FE models and seven tested specimens. With 10 input parameters for the neural networks, the response of SMA-based connections was characterized using seven parameters, including rotation (θB, θC, θE), moment (MB, MC, ME), and self-centering response parameter (β). For each of these response parameters, a neural network was trained. The trained neural networks were then used to develop a GUI for predicting the response of SMA-based endplate connections. The results demonstrated the high prediction accuracy of the proposed model for each response.
The predictive models were then used to perform two optimization studies based on a multiobjective genetic algorithm. The optimization objectives included minimizing the material use and improving the connection response characteristics. For each optimization problem, a set of trade-off optimal solutions (Pareto front) were determined. A desirability approach was then used to select the optimal solutions with the highest desirability scores. The predictive models were also used to develop a phenomenological model for SMA connections in OpenSees. The use of the predictive tool was illustrated by developing OpenSees models and performing pushover and response history analyses on SMA connections and self-centering frames. The following conclusions are summarized:
The proposed ANNs can accurately predict the response of extended endplate connections.
The developed MATLAB tool can accurately and efficiently predict the backbone and self-centering response of SMA-based extended endplate connections. When aiming to determine the moment-rotation response of the connection only, this tool may eliminate the need for a detailed FE analysis of SMA connections.
By using the proposed phenomenological model together with the developed predictive tool, computationally efficient OpenSees models can be developed.
Larger bolt diameters and higher maximum transformation strains are required for SMA bolts to increase the moment and rotation capacity of SMA connections. The optimization results demonstrate that higher bolt diameter and beam depth increase the initial stiffness and moment capacity. Further, it is shown that longer bolts can improve the rotational capacity of SMA-based connections.
Pushover results show that MRFs equipped with optimized SMA-based connections demonstrate greater lateral load resistance and drift capacity.
Nonlinear response history analysis results show that MRFs with optimized SMA-based connections can exhibit negligible residual deformations.
The developed predictive tool should be used only for endplate connections with SMA bolts and only with the factor ranges in Table 7. Furthermore, the SMA connection is assumed to be designed with a thick endplate to confine large deformations into the SMA bolts. Beams and columns should be checked to meet the requirements of AISC 341-16 (AISC 2016b) for highly ductile sections. The AISC 358-16 (AISC 2016a) requirements for extended endplate connections should also be checked, including the requirements for the beam and column sizes, beam clear span-to-depth ratio, beam and column lateral bracing, and strong-column-weak-beam. Additionally, the self-centering response prediction in this study neglects small residual deformations of SMA materials.
Open-source platforms (e.g., Jupyter notebook) could be used in future research for developing predictive models. Moreover, the developed ANNs and predictive tools in this paper do not account for different sources of uncertainty, such as the randomness of measurement and material. Future research is recommended to quantify uncertainties and evaluate their effects on response predictions developed based on deterministic approaches. In the case of finding significant variations, attempts can be made to apply adjustment or safety factors to the response values predicted using a deterministic approach. Further research can finetune other ML algorithms and possibly improve the accuracy of the developed predictive tool.

Data Availability Statement

The data, models, or code that support the findings of this study are available from the corresponding author upon reasonable request. The predictive tool generated in the study is publicly available in a Zenodo repository (Mohammadi Nia and Moradi 2022) in accordance with funder data retention policies.

Acknowledgments

The financial contributions of the Toronto Metropolitan University Faculty of Engineering and Architectural Science and the Natural Sciences and Engineering Research Council of Canada (NSERC) through the Discovery Grant are gratefully acknowledged.

References

Abdalla, K. M., and G. E. Stavroulakis. 1995. “A backpropagation neural network model for semi-rigid steel connections.” Comput. Civ. Infrastruct. Eng. 10 (2): 77–87. https://doi.org/10.1111/j.1467-8667.1995.tb00271.x.
AISC. 2016a. Prequalified connections for special and intermediate steel moment frames for seismic applications. ANSI/AISC 358-16. Chicago: AISC.
AISC. 2016b. Seismic provisions for structural steel buildings. ANSI/AISC 341-16. Chicago: AISC.
Anderson, D., E. L. Hines, S. J. Arthur, and E. L. Eiap. 1997. “Application of artificial neural networks to the prediction of minor axis steel connections.” Comput. Struct. 63 (4): 685–692. https://doi.org/10.1016/S0045-7949(96)00080-6.
ANSYS. 2020. Mechanical APDL (release 2020 R1). Canonsburg, PA: ANSYS.
ASCE. 2017. Seismic evaluation and retrofit of existing buildings. Reston, VA: ASCE.
Caglar, N. 2009. “Neural network based approach for determining the shear strength of circular reinforced concrete columns.” Constr. Build. Mater. 23 (10): 3225–3232. https://doi.org/10.1016/j.conbuildmat.2009.06.002.
Deb, K. 2011. “Multi-objective optimisation using evolutionary algorithms: An introduction.” In Multi-objective evol. optim. prod. des. manuf., 3–34. London: Springer.
Degtyarev, V. V., and M. Z. Naser. 2021. “Boosting machines for predicting shear strength of CFS channels with staggered web perforations.” Structures 34 (Oct): 3391–3403. https://doi.org/10.1016/j.istruc.2021.09.060.
Fang, C., M. C. H. Yam, A. C. C. Lam, and L. Xie. 2013. “Cyclic performance of extended end-plate connections equipped with shape memory alloy bolts.” J. Constr. Steel Res. 94 (Dec): 122–136. https://doi.org/10.1016/j.jcsr.2013.11.008.
Fausett, L. V. 2006. Fundamentals of neural networks: Architectures, algorithms and applications. London: Pearson Education.
Hammoudi, A., K. Moussaceb, C. Belebchouche, and F. Dahmoune. 2019. “Comparison of artificial neural network (ANN) and response surface methodology (RSM) prediction in compressive strength of recycled concrete aggregates.” Constr. Build. Mater. 209 (Mar): 425–436. https://doi.org/10.1016/j.conbuildmat.2019.03.119.
Horton, T. A., I. Hajirasouliha, B. Davison, and Z. Ozdemir. 2021. “Accurate prediction of cyclic hysteresis behaviour of RBS connections using deep learning neural networks.” Eng. Struct. 247 (Sept): 113156. https://doi.org/10.1016/j.engstruct.2021.113156.
Huang, H., and H. V. Burton. 2019. “Classification of in-plane failure modes for reinforced concrete frames with infills using machine learning.” J. Build. Eng. 25 (2): 100767. https://doi.org/10.1016/j.jobe.2019.100767.
Kim, S. H., X. Song, C. Cho, and C. H. Lee. 2021. “Strength prediction of steel CHS X-joints via leveraging finite element method and machine learning solutions.” J. Constr. Steel Res. 176 (12): 106394. https://doi.org/10.1016/j.jcsr.2020.106394.
Ma, H., C. Cho, and T. Wilkinson. 2008. “A numerical study on bolted end-plate connection using shape memory alloys.” Mater. Struct. 41 (8): 1419–1426. https://doi.org/10.1617/s11527-007-9339-5.
Ma, H., T. Wilkinson, and C. Cho. 2007. “Feasibility study on a self-centering beam-to-column connection by using the superelastic behavior of SMAs.” Smart Mater. Struct. 16 (5): 1555–1563. https://doi.org/10.1088/0964-1726/16/5/008.
Majidifard, H., B. Jahangiri, W. G. Buttlar, and A. H. Alavi. 2019. “New machine learning-based prediction models for fracture energy of asphalt mixtures.” Measurement 135 (Dec): 438–451. https://doi.org/10.1016/j.measurement.2018.11.081.
Mohammadi Nia, M., and S. Moradi. 2020. “Limit state behavior and response sensitivity analysis of endplate steel connections with shape memory alloy bolts.” J. Intell. Mater. Syst. Struct. 31 (18): 2071–2087. https://doi.org/10.1177/1045389X20942584.
Mohammadi Nia, M., and S. Moradi. 2021. “Surrogate models for endplate beam-column connections with shape memory alloy bolts.” J. Constr. Steel Res. 187 (Sep): 106929. https://doi.org/10.1016/j.jcsr.2021.106929.
Mohammadi Nia, M., and S. Moradi. 2022. SMA_Conn_ResModel: First release, v 1.0.0. Geneva: Zenodo.
Montgomery, D. C. 2006. “Design and analysis of experiments.” Technometrics 48 (1): 158. https://doi.org/10.1198/tech.2006.s372.
Moradi, S., and M. S. Alam. 2015a. “Feasibility study of utilizing superelastic shape memory alloy plates in steel beam–column connections for improved seismic performance.” J. Intell. Mater. Syst. Struct. 26 (4): 463–475. https://doi.org/10.1177/1045389X14529032.
Moradi, S., and M. S. Alam. 2015b. “Finite-element simulation of posttensioned steel connections with bolted angles under cyclic loading.” J. Struct. Eng. 142 (1): 04015075. https://doi.org/10.1061/(asce)st.1943-541x.0001336.
Naderpour, H., A. H. Rafiean, and P. Fakharian. 2018. “Compressive strength prediction of environmentally friendly concrete using artificial neural networks.” J. Build. Eng. 16 (Jan): 213–219. https://doi.org/10.1016/j.jobe.2018.01.007.
Naser, M., G. Abu-Lebdeh, and R. Hawileh. 2012. “Analysis of RC T-beams strengthened with CFRP plates under fire loading using ANN.” Constr. Build. Mater. 37 (Jun): 301–309. https://doi.org/10.1016/j.conbuildmat.2012.07.001.
Naser, M. Z., and A. H. Alavi. 2021. “Error metrics and performance fitness indicators for artificial intelligence and machine learning in engineering and sciences.” Archit. Struct. Constr. 2021 (Jan): 1–19. https://doi.org/10.1007/s44150-021-00015-8.
Naser, M. Z., V. Kodur, H. T. Thai, R. Hawileh, J. Abdalla, and V. V. Degtyarev. 2021. “StructuresNet and FireNet: Benchmarking databases and machine learning algorithms in structural and fire engineering domains.” J. Build. Eng. 44 (Jul): 102977. https://doi.org/10.1016/j.jobe.2021.102977.
Ngatchou, P., A. Zarei, and M. A. El-Sharkawi. 2005. “Pareto multi objective optimization.” In Proc., 13th Int. Conf. Intelligent System Application to Power System ISAP’05, 84–91. New York: IEEE.
OpenSees. 2013. Open system for earthquake engineering simulation (OpenSees), v 3.3.0 [Computer Software]. Berkeley, CA: Pacific Earthquake Engineering Research Center.
Qiu, C.-X., and S. Zhu. 2017. “Performance-based seismic design of self-centering steel frames with SMA-based braces.” Eng. Struct. 130 (Sep): 67–82. https://doi.org/10.1016/j.engstruct.2016.09.051.
Sabouri Ghannad, M. M., M. Mohammadi Nia, and S. Moradi. 2021. “Predictive equations and continuum FE models for self-centering response of SMA-based steel beam-column connections with steel angles.” J. Build. Eng. 43 (Nov): 102859. https://doi.org/10.1016/j.jobe.2021.102859.
Salehi, H., and R. Burgueño. 2018. “Emerging artificial intelligence methods in structural engineering.” Eng. Struct. 171 (Nov): 170–189. https://doi.org/10.1016/j.engstruct.2018.05.084.
Sediek, O. A., T. Wu, J. McCormick, and S. El-Tawil. 2022. “Prediction of seismic collapse behavior of deep steel columns using machine learning.” Structures 40 (Oct): 163–175. https://doi.org/10.1016/j.istruc.2022.04.021.
Shah, S. N. R., N. H. Ramli Sulong, and A. El-Shafie. 2018. “New approach for developing soft computational prediction models for moment and rotation of boltless steel connections.” Thin-Walled Struct. 133 (12): 206–215. https://doi.org/10.1016/j.tws.2018.09.032.
Smucker, B., M. Krzywinski, and N. Altman. 2018. “Optimal experimental design.” Nat. Methods 15 (8): 559–560. https://doi.org/10.1038/s41592-018-0083-2.
Somerville, P. G. 1997. Development of ground motion time histories for phase 2 of the FEMA/SAC steel project. Washington, DC: SAC Joint Venture.
Speicher, M. S., R. DesRoches, and R. T. Leon. 2011. “Experimental results of a NiTi shape memory alloy (SMA)-based recentering beam-column connection.” Eng. Struct. 33 (9): 2448–2457. https://doi.org/10.1016/j.engstruct.2011.04.018.
Stavroulakis, G. E., A. V. Avdelas, K. M. Abdalla, and P. D. Panagiotopoulos. 1997. “A neural network approach to the modelling, calculation and identification of semi-rigid connections in steel structures.” J. Constr. Steel Res. 44 (1–2): 91–105. https://doi.org/10.1016/S0143-974X(97)00039-4.
Szandała, T. 2021. “Review and comparison of commonly used activation functions for deep neural networks.” Stud. Comput. Intell. 903 (11): 203–224. https://doi.org/10.1007/978-981-15-5495-7_11.
Wang, W., C. Fang, and J. Liu. 2017. “Self-centering beam-to-column connections with combined superelastic SMA bolts and steel angles.” J. Struct. Eng. 143 (2): 04016175. https://doi.org/10.1061/(ASCE)ST.1943-541X.0001675.
Wilson, J. C., and M. J. Wesolowsky. 2005. “Shape memory alloys for seismic response modification: A state-of-the-art review.” Earthquake Spectra 21 (2): 569–601. https://doi.org/10.1193/1.1897384.
Xu, J., J. Zhao, Z. Song, and M. Liu. 2012. “Prediction of ultimate bearing capacity of tubular T-joint under fire using artificial neural networks.” Saf. Sci. 50 (7): 1495–1501. https://doi.org/10.1016/j.ssci.2012.02.004.
Yam, M. C. H., C. Fang, A. C. C. Lam, and Y. Zhang. 2015. “Numerical study and practical design of beam-to-column connections with shape memory alloys.” J. Constr. Steel Res. 104 (10): 177–192. https://doi.org/10.1016/j.jcsr.2014.10.017.
Yaro, N. S. A., M. H. Sutanto, N. Z. Habib, M. Napiah, A. Usman, and A. Muhammad. 2022. “Comparison of response surface methodology and artificial neural network approach in predicting the performance and properties of palm oil clinker fine modified asphalt mixtures.” Constr. Build. Mater. 324 (22): 126618. https://doi.org/10.1016/j.conbuildmat.2022.126618.
Yun, G. J., J. Ghaboussi, and A. S. Elnashai. 2008. “A design-variable-based inelastic hysteretic model for beam–column connections.” Earthquake Eng. Struct. Dyn. 37 (4): 535–555. https://doi.org/10.1002/eqe.771.
Zarringol, M., H. T. Thai, and M. Z. Naser. 2021. “Application of machine learning models for designing CFCFST columns.” J. Constr. Steel Res., 185 (Jul): 106856. https://doi.org/10.1016/j.jcsr.2021.106856.

Information & Authors

Information

Published In

Go to Journal of Structural Engineering
Journal of Structural Engineering
Volume 148Issue 12December 2022

History

Received: Feb 20, 2022
Accepted: Jun 22, 2022
Published online: Sep 28, 2022
Published in print: Dec 1, 2022
Discussion open until: Feb 28, 2023

Authors

Affiliations

Majid Mohammadi Nia, S.M.ASCE [email protected]
Ph.D. Candidate, Dept. of Civil Engineering, Toronto Metropolitan Univ., Toronto, ON, Canada M5B 2K3. Email: [email protected]
Assistant Professor, Dept. of Civil Engineering, Toronto Metropolitan Univ., Toronto, ON, Canada M5B 2K3 (corresponding author). ORCID: https://orcid.org/0000-0002-5075-3110. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share