Next Article in Journal
In-Plane Mechanical Properties Test of Prefabricated Composite Wall with Light Steel and Tailings Microcrystalline Foamed Plate
Previous Article in Journal
Development of Post-Industrial Heritage Landscape Design Based on Visual Cognitive Schema Theory: A Case Study of the Shou Gang Industrial Cultural Heritage Site
Previous Article in Special Issue
Performance Assessment of Wood-Based Composite Materials Subjected to High Temperatures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Risk Assessment of Overturning of Freestanding Non-Structural Building Contents in Buckling-Restrained Braced Frames

1
Graduate School of Engineering, Tohoku University, Sendai 980-8579, Japan
2
International Research Institute of Disaster Science, Tohoku University, Sendai 980-8577, Japan
*
Author to whom correspondence should be addressed.
Buildings 2024, 14(10), 3195; https://doi.org/10.3390/buildings14103195
Submission received: 8 September 2024 / Revised: 1 October 2024 / Accepted: 5 October 2024 / Published: 8 October 2024

Abstract

:
The increasing demand in structural engineering now extends beyond collapse prevention to encompass business continuity planning (BCP). In response, energy dissipation devices have garnered significant attention for building response control. Among these, buckling-restrained braces (BRBs) are particularly favored due to their stable hysteretic behavior and well-established design provisions. However, BCP also necessitates the prevention of furniture overturning—an area that remains quantitatively underexplored in the context of buckling-restrained braced frames (BRBFs). Addressing this gap, this research designs BRBFs using various design criteria and performs incremental dynamic analysis (IDA) with artificially generated seismic waves. The results are compared with previously developed fragility curves for furniture overturning under different BRB design conditions. The findings demonstrate that the fragility of furniture overturning can be mitigated by a natural frequency shift, which alters the threshold of critical peak floor acceleration. These results, combined with hazard curves obtained from various locations across Japan, quantify the mean annual frequency of furniture overturning. The study reveals that increased floor acceleration in stiffer BRBFs can lead to a 3.8-fold higher risk of furniture overturning compared to frames without BRBs. This heightened risk also arises from the greater hazards at shorter natural periods due to stricter response reduction demands. The probabilistic risk analysis, which integrates fragility and hazard assessments, provides deeper insights into the evaluation of BCP.

1. Introduction

In the face of significant earthquakes occurring globally, appropriate seismic retrofitting is essential. One effective method for collapse prevention is the implementation of seismic retrofitting or strengthening by buckling-restrained braces (BRBs). A BRB consists of a steel core encased in concrete, which is further confined by a steel tube. Typically, BRBs are installed in diagonal, V, or chevron (inverted-V) configurations, as depicted in Figure 1.
Since the invention of the BRB by Kimura et al. [1], extensive research has been conducted to evaluate BRBs at the component level [2,3,4]. These studies have demonstrated that BRBs possess exceptional ductility and energy-dissipation capacity, highlighting their ability to withstand strong earthquakes.
Currently, the design procedure is outlined in the prevailing design specifications [5,6]. The method for calculating the required amount of BRBs was investigated by Kasai et al. [7], and the proposed method is now incorporated into the guidelines of the Architectural Institute of Japan (AIJ) [5]. The necessary design considerations for the main frame have been extensively studied at both the member and frame levels [8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31].
Thanks to these major advancements, buildings equipped with BRBs can withstand more significant earthquakes. As illustrated in Figure 2a,b, buckling-restrained braced frames (BRBFs) effectively prevent collapse even during large-scale seismic events, which could otherwise lead to the collapse of moment-resisting frames (MRFs). In recent years, the structural engineering field has increasingly emphasized the importance of ensuring business continuity after a major earthquake, aligning with the concept of Business Continuity Planning (BCP). However, previous research has predominantly focused on collapse prevention [32,33,34,35]. Faced with this need, structural health monitoring and damage detection methods have been emerging and have recently advanced [36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51]. While they effectively contribute to resilience, avoiding any damage that impacts BCP is the best option. Figure 2c illustrates the interior of a room following the 2022 Fukushima-Oki earthquake. The building was equipped with an energy dissipation device. Although the building remained intact and did not collapse, furniture toppled over, leading to a suspension of business operations during the repair work. Particularly after the Great East Japan Earthquake (GEJE) in 2011, researchers have shown increased interest in the damage to non-structural components, such as ceiling failures [52,53].
Regarding furniture overturning, a fundamental study was conducted by Kaneko [54,55]. The essential characteristics of overturning ratios of rigid bodies subjected to large input motions were examined through analytical studies. The results revealed that overturning ratios are influenced by the sizes and shapes of rigid bodies, as well as the levels and predominant frequencies of the input motions. Based on these studies, a general fragility curve was proposed to describe the relationship between overturning ratios and the amplification of input motions.
Subsequent studies have advanced the understanding of the fragility of rocking bodies by considering deformation, nonlinearity, and sliding [56,57,58,59,60,61]. These studies typically derived the moment equilibrium around the center of rotation using the moment of inertia and applied horizontal forces. Additionally, deformation is modeled using a spring and dashpot system at the corners of the rocking bodies to better reflect realistic conditions.
Subsequently, Saito et al. [62] conducted a seismic risk evaluation study considering furniture overturning in mid- and low-rise office buildings. Saito et al. [62] focused on mid- and low-rise steel frame office buildings. The study used the ratio of Qu/Qun (the horizontal load-bearing capacity, Qu, to the required horizontal yield strength, Qun) as an analytical parameter. Seismic response analysis was used to evaluate response values, and building damage was assessed through Seismic Risk Analysis. Additionally, the impact of furniture within the building was analyzed by determining the overturning ratio. The study demonstrated that by increasing Qu/Qun, the accuracy of seismic safety analysis for the building can be enhanced.
The probabilistic risk assessment (PRA) framework is well-suited for quantifying the effectiveness of seismic retrofitting in preventing furniture overturning. In a foundational study, Ishikawa et al. [63] proposed a procedure for seismic risk evaluation in buildings, incorporating the results of probabilistic seismic hazard assessment (PSHA), structural fragility analysis, and economic loss estimation. The study produced a seismic risk curve that illustrates the relationship between financial loss and its annual exceedance probability, with case studies validating the proposed methodology. This approach is widely applied globally [64,65,66,67].
PRA typically involves incremental dynamic analysis (IDA), comprehensively outlined by Vamvatsikos and Cornell [68]. IDA is a parametric method developed in various forms to estimate structural performance under seismic loads. It entails applying scaled ground motion records to a structural model across multiple intensity levels, producing response curves parameterized by intensity. Their study established unified terminology, introduced suitable algorithms, and examined the properties of IDA curves for both single- and multi-degree-of-freedom systems. Additionally, it discussed methods for summarizing multi-record IDA results, comparing IDA with conventional static pushover analysis, and evaluating the yield reduction R-factor.
Eads et al. [69] investigated various aspects of calculating the mean annual frequency of collapse and introduced an efficient approach for estimating the sideways collapse risk of structures in seismic regions. The deaggregation of the mean annual frequency of collapse showed that this risk is generally dominated by earthquake ground motion intensities in the lower half of the collapse fragility curve. Their study also quantified the uncertainty in both the collapse fragility curve and the mean annual frequency of collapse based on the number of ground motions used in the analysis. The proposed method significantly reduced the computational effort and the uncertainty associated with these estimates.
Deylami and Mahdavipour [70] critiqued a significant drawback of BRBFs, namely the low post-yield stiffness of steel cores, which results in large residual drifts in a story after earthquakes. Their study examined the seismic demands of low- and mid-rise BRBFs and Dual-BRBFs using Probabilistic Seismic Demand Analysis (PSDA). By comparing demand hazard curves, they concluded that employing BRBFs as a dual system could significantly reduce residual drift demands, improving the resilience of such structures after earthquakes with lower repair costs. Additionally, the study used two nonlinear models—a deteriorating model and a non-deteriorating model—to examine the impact of degradation on Dual-BRBFs’ seismic demands. Their analysis revealed residual deformations are more susceptible to degradation than maximum deformations. However, deterioration is mitigated by the significant stiffness of BRBs, which keeps MRFs within a low range of nonlinearity.
The aforementioned studies primarily concentrated on structural collapse or damage related to structural safety. However, as previously noted, the recent societal demand increasingly emphasizes the importance of maintaining business continuity even after major earthquakes. Despite this, approaches to BCP based on PRA remain limited. In response to this, the present study specifically focuses on applying PRA to the issue of furniture overturning in BRBFs. Additionally, the effectiveness of retrofitting with BRBs for BCP is evaluated analytically. Therefore, building collapse is not assessed in this study, as advancements in BRBF design have already enhanced the structural capacity to prevent collapse effectively.
In this study, the model structure is derived from the design provisions of the AIJ and retrofitted with BRBs according to various design criteria. Realistic seismic waves are generated, reflecting the underlying ground structure and potential fault fracture scenarios. Using these seismic waves, IDA is conducted to establish fragility curves for furniture overturning. Finally, fragility and hazard curves for various locations in Japan are compared, and the mean annual frequency of furniture overturning is computed.
The findings of this research contribute to advancing seismic design by incorporating considerations for BCP. Moreover, the study highlights the importance of adequately securing furniture, even in buildings retrofitted with BRBs. In practical applications, engineers can quantify the risk of furniture overturning and assess its impact on BCP, providing a reasonable criterion for discussions with building owners or tenants. As such, the methodology presented in this research offers a platform for developing more sophisticated design strategies that account for business contingency.

2. Model Structure for Seismic Retrofitting Using Buckling-Restrained Braces

2.1. Outline of MRF Model Structure

This study considers the seismic upgrading of a four-story Japanese office building. The structure in question is a representative model of a steel moment-resisting frame (MRF) as specified in the provisions of the AIJ [5].
The structure represents a typical office building designed in accordance with Japanese regulations, featuring a regular floor plan. Consequently, torsional response is not a significant concern. This regularity makes the building suitable for studying the relationship between retrofitting with BRBs and the risk of furniture overturning.
A drawing of the structure is presented in Figure 3. The building’s floor plan spans 25.6 m in the x direction and 19.2 m in the y direction. The weight distribution for each story is shown in Figure 3. The columns are square hollow section (SHS) tubes, with widths ranging from 350 to 450 mm and wall thicknesses between 16 and 22 mm.
The main girders consist of wide-flange sections with depths ranging from 550 to 800 mm and flange widths of 200 to 300 mm, with plate thicknesses between 9 and 25 mm. All girder-column connections are designed as rigid moment connections. The steel used is SN490, per Japanese standards [71], with a yield stress of 325 N/mm2.

2.2. Equivalent Linearization-Based Design of a Buckling-Restrained Brace

For this study, the BRBs are designed based on the principles of equivalent linearization, a method introduced initially by Kasai et al. [7] and currently codified in the Japanese design specifications [5]. Figure 4 illustrates the concept of response reduction of BRBFs. Respective figures indicate the idealized displacement, velocity, and acceleration spectrum. The design primarily aims to reduce the displacement response. This stems from the enhanced stiffness using the added BRBs. Also, the enlargement of the damping factor reduces the response spectrum; thereby, the displacement response further decreases, as presented in Figure 4a.
Using mathematical expressions for the effective vibration period and damping ratio of a building equipped with BRBs, along with idealized seismic response spectra, the peak seismic responses of the system and its local components can be expressed as continuous functions of structural and seismic parameters. These relationships are depicted in “performance curves”.
Figure 5 illustrates the concept of equivalent linearization for a BRB integrated with an elastic frame, represented as an equivalent single-degree-of-freedom (SDOF) system. Figure 5a shows the displacement response spectrum used in this study, where the vertical axis represents the displacement response, denoted as Se/ω2, and the horizontal axis represents the natural period, T. The design spectrum used in this study is derived from Eurocode-8 [72], with the ground type assumed to be Type C and the reference ground acceleration, ag, set at 3.0 m/s2. Additionally, importance factors γI of 1.0, 1.2, and 1.4, corresponding to importance classes II, III, and IV, respectively, are applied in the BRB design.
Five scenarios are considered for seismic upgrading:
  • θt = 1/120 rad with γIag = 3.0 m/s2 peak ground acceleration (PGA),
  • θt = 1/150 rad with γIag = 3.0 m/s2 PGA,
  • θt = 1/200 rad with γIag = 3.0 m/s2 PGA,
  • θt = 1/200 rad with γIag = 3.6 m/s2 PGA, and
  • θt = 1/200 rad with γIag = 4.2 m/s2 PGA.
Here, θt represents the target story drift.
Figure 5b illustrates the performance curve, which consists of the displacement reduction ratio (Rd) and the force (or pseudo-acceleration) reduction ratio (Rpa), both expressed relative to the frame’s responses without BRBs. In equivalent linearization, the target story drift θt is initially determined as a sufficiently small value to prevent structural damage. The displacement response can be computed using the equivalent SDOF system and response spectra. This response is converted into story drift θ by dividing by the effective height, hef, of the equivalent SDOF. The displacement reduction ratio, Rd, is calculated as Rd = θt/θ.
The curve depicted in Figure 5b theoretically captures the influence of the stiffness ratio Ka/Kf between the added component (a damper and brace combined in series) and the frame, as well as the ductility demand μ of the added component. These curves incorporate these critical factors and provide the necessary Ka/Kf and μ values to achieve the target drift or displacement reduction ratio Rd. The required yield force of the damper can also be derived from the Rd and force reduction ratio Ra values. The results of this study suggest that the design ductility of the BRB is 4.0, as the performance curve identifies this value as the optimal response reduction point, indicated by the convex portion in Figure 5b.
Once the necessary Ka/Kf value is determined for the SDOF system, it can be translated into the requirements for dampers in a multi-story system under the following assumptions:
  • The equivalent damping factor is consistent between the SDOF and multi-story systems.
  • The inter-story drift is uniform across all stories and equals the target story drift for the seismic force, calibrated according to the Ai distribution.
  • The ductility of the BRB is uniform across all stories.
Here, the Ai distribution represents the vertical seismic force distribution guided by the AIJ [73]. The necessary stiffness of individual stories Kai is computed using the following equations to meet these criteria.
K a i = Q i h i i = 1 N K f i h i 2 i = 1 N Q i h i μ + K a K f μ · K f i
In this context, Qi represents the story shear force, hi denotes the height of the i-th floor, Kfi signifies the story stiffness of the i-th floor, and μ denotes the ductility of the BRBs at the target story drift. Additionally, the required yield axial displacement uayi is determined based on the design ductility μ, the story height hi, and the target story drift θt.
u a y i = θ t μ · h i
Ultimately, the yield axial forces of the respective BRBs, Fay, can be computed based on the axial stiffness Kai and the yield axial displacement uay. The formulation for Fay is provided below.
F a y i = K a i · u a y i
The calibrated specifications for the BRBs are detailed in Table 1. These BRBs are positioned in bay Y2, which is situated at the center of the frame. The design-based shear ratios are 0.47, 0.57, and 0.66 for seismic accelerations of γIag = 3.0 m/s2, 3.6 m/s2, and 4.2 m/s2, respectively.

3. Outline of Nonlinear Response History Analyses and Fragility Assessment of Furniture Overturning

3.1. Outline of Nonlinear Response History Analyses

A three-dimensional model was constructed, as detailed in Figure 6. Nonlinear response history analyses (NRHA) were conducted using ABAQUS ver. 2021, a finite element analysis (FEA) software package. For specific settings and element definitions, readers are referred to the ABAQUS manual [74]. The ABAQUS/Standard module is employed for the analyses described. The model utilizes a three-dimensional configuration with encastre constraints applied at the column base. As this research focuses on behavior along the x-axis, out-of-plane (OOP) deformation is constrained by the boundary conditions.
Beams and columns are represented using beam elements (B31), with distributed plasticity modeling adopted. Member yielding is identified based on von Mises stress, considering the sum of stress from both axial force and flexural moment for yield detection. The yield strength is determined to be 325 N/mm2. Elastic-perfectly plastic (EPP) hysteresis is applied to each member, with column and beam intersections defined as rigid bodies.
The flexural stiffness of girders is doubled for beams with concrete slabs on both sides and increased by 1.5 times for beams with a half slab in order to model the influence of the concrete slab. The contributions of the concrete slabs to flexural stiffness are calculated according to Japanese design guidelines [75]. The contribution of the concrete slab is influenced by the shear connector performance [76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112]. Also, structural members can originate the buckling in a huge deformation, as summarized in the previous studies [113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139]. However, this research assumes idealized conditions to examine the risk of furniture overturning specifically. Additionally, Rayleigh damping is utilized for each structural member, with primary and secondary damping ratios set at 0.02.
The established model was verified by comparing the natural period and load-displacement relationship with those provided in the AIJ provision [5].

3.2. Fragility Assessment of Furniture Overturning

The fragility of furniture overturning is assessed statistically. Although various approaches exist for evaluating fragility [56,57,58,59,60,61], this research adopts the methodology proposed by Kaneko [54]. Kaneko et al. [54] integrated a theoretical framework with post-earthquake inspections of overturned bodies following several large seismic events. The resulting trend is represented by a continuous function, making it straightforward to apply when determining the threshold acceleration. The application of alternative evaluation methods will be explored in future research.
Based on a previous study [54], the probability of furniture overturning is determined using the following equations. The acceleration that results in a 50% rate of furniture overturning is calculated using Equation (6).
R A f = α · φ l n A f l n A R 50 ζ
A R 50 = B H · g 1 + B H F f F b 10 B H 1 + B H 2.5 · 2 π F f F f > F b
F f = V f 2 π D f
where φ represents the log-normal distribution with a mean of lnAR50 and a standard deviation of ζ (where ζ = 0.50). Af denotes the peak floor acceleration (PFA) response, Vf the peak floor response velocity, Df the peak floor response displacement, α is the coefficient representing the influence of slip, B is the depth of the furniture, H is the height of the furniture, and g is the gravitational acceleration. In this research, the constants in Equation (5) are determined as α = 0.8 and ζ = 0.50.
This study examines three types of furniture with dimensions specified in Table 2. The depth of furniture B remains constant while the height H is varied to modify the dimensions. The respective cases are categorized as “low”, “middle”, and “high”. The furniture is assumed to be not anchored to the walls or floors. The frequency of furniture Fb varies depending on the height of the furniture.
F b = 15.6 H 1 + B H 1.5
The fragility curves calculated using the above equations are presented in Figure 7. In Figure 7, the taller furniture is more likely to overturn at lower floor accelerations.

4. Computation of Artificial Seismic Waves Reflecting Ground Characteristics

The seismic waves are calculated considering the rupture process of faults and the ground structure. Three locations in central Sendai were selected, focusing on the Nagamachi-Rifu Fault (M7.5) located directly beneath the city. Figure 8 illustrates the surface projection of the characterized source model from the National Seismic Hazard Map for Japan [140] and the locations of the two strong motion generation areas [141]. In this analysis, two cases with different rupture initiation points are evaluated. The respective locations are marked by ★.
The evaluation of seismic waveforms follows the methodology outlined in the National Seismic Hazard Map for Japan [140]. First, the waveforms at the engineering bedrock level are predicted using the stochastic waveform synthesis method [142]. The component waveforms are artificially generated using the envelope shape from Boore [143] and random phase. The Q value of the propagation path is given by the following equation [144]:
Q = 110 f 0.7
where f represents the frequency. The amplification rate of the deep ground from the seismic bedrock to the engineering bedrock is derived from the 1D structure directly beneath each site, using the subsurface structure model applied in earthquake damage estimation for Sendai [145]. Ultimately, the amplification rate is calibrated under the condition of vertical incidence of S-waves. The response of the surface ground layer above the engineering bedrock is approximated using a modified R-O model [146], which accounts for strain-dependent rigidity and attenuation as used in the earthquake damage simulation for Sendai [145]. Surface acceleration waveforms are then obtained through a fully nonlinear time-domain analysis of total stress, using the seismic motion at the engineering bedrock as the input.
The coordinates of the three locations, along with the PGA and PGV of the calculated waveforms at the surface, are shown in Table 3. Figure 9 presents the acceleration response spectra with a damping factor of 2%. All evaluation points are located in central Sendai City, near the upper edge of the strong motion generation area on the southern side. The differences observed between the evaluation points are primarily due to variations in ground conditions.

5. Contribution of Seismic Retrofitting to Structural Response Reduction Based on Incremental Dynamic Analysis

The input is standardized based on the acceleration response of the Single-Degree-of-Freedom (SDOF) system with a damping factor of 0.02. The natural periods of the structure for the MRF and BRBFs are derived from eigenvalue analysis. The intervals range from 0.1 G to 3.0 G with a step size of 0.1 G, resulting in a total of 2160 cases. The seismic waves utilized in the IDA are the 12 waves generated in the preceding chapter.
Figure 10 presents the peak story drift for the MRF and BRBFs. The horizontal axis represents Sa(T), and the vertical axis shows the peak story drift. Since the story drift becomes the most significant in the first story, the floor displacement of the second story δ2 is extracted in Figure 10. The median and 16th/84th percentile lines are depicted by bold black lines. Overall, the story drift decreases with more stringent design criteria, and building collapse can be prevented by retrofitting with BRBFs.
Figure 11 summarizes the PFA. According to previous studies [54], furniture overturning is associated with PFA. Unlike story drift, PFA cannot be mitigated by more stringent BRB design criteria (see Figure 4c). Increased BRB stiffness, intended to reduce displacement, may result in higher acceleration responses.
Additionally, the occurrence of furniture overturning is related to the building’s equivalent frequency, as considered by Equation (7). The AR50 differs depending on the balance of the unique frequency of furniture Fb and the characteristic frequency of the building, thereby altering the fragility curve. The threshold acceleration for 50% of furniture overturning is shown in Figure 12. The furniture dimension is low. The median and 16th/84th percentile lines are also depicted in the figure. The equivalent frequency is calculated based on the building’s response, precisely the peak floor velocity Vf and peak floor displacement Df, with the equation provided in Equation (7). Figure 12 demonstrates that the threshold acceleration increases with more stringent BRB design criteria primarily due to smaller peak floor displacements.
Figure 13 illustrates the calculation flow for deriving the fragility curve of furniture overturning. As summarized in Figure 11, the PFA is determined for each story. Simultaneously, the threshold acceleration for a 50% probability of furniture overturning is calculated using the equation provided by Kaneko [54]. The exceedance of this threshold is then evaluated for each ground motion. Finally, the exceedance ratio is computed for the respective spectral accelerations. The fragility curve is obtained by fitting a cumulative distribution function to a log-normal distribution.
Figure 14 illustrates the probability of furniture overturning. The horizontal axis represents the intensity of the input wave, and the vertical axis denotes the probability of exceeding 50% furniture overturning. The plots suggest that the likelihood of overturning decreases with more stringent design criteria. Although PFA can increase with greater BRB stiffness and capacity, the threshold acceleration also rises with more stringent BRB design criteria. Consequently, the fragility curve for θt = 1/200 rad (γIag = 4.2 m/s2) becomes the smallest in Figure 14.

6. Evaluation of Mean Annual Frequency of Furniture Overturning

To assess the effectiveness of seismic retrofitting in business continuity planning, the mean annual frequency of furniture overturning is determined based on the framework depicted in Figure 15.
This study utilizes the mean annual frequency λc for risk evaluation. Calculating λc requires two key elements: the seismic hazard curve, which represents the mean annual frequency of exceeding ground motion intensities at a specific site, and the fragility curve for furniture overturning. Ground motion intensity is quantified using an intensity measure (IM), such as Sa(T), the spectral acceleration at the structure’s fundamental period.
The value of λc is obtained by integrating the furniture fragility curve with the site-specific seismic hazard curve using the following equation:
λ c = 0 P C | i m · d λ I M i m
where P(CIM) represents the probability of furniture overturning under an earthquake with a given ground motion intensity level IM, and λIM denotes the mean annual frequency of exceeding the ground motion intensity IM. By multiplying and dividing the right-hand side of Equation (10) by d(IM), the expression for calculating λc can be reformulated as the following:
λ c = 0 P C | i m · d λ I M i m d i m d i m
where IM(IM)/d(IM) represents the slope of the seismic hazard curve at the site. A closed-form solution for the integral in Equation (11) is usually unavailable, so the integral is typically evaluated through numerical integration. This involves computing the product of the collapse probability conditioned on IM, the slope of the seismic hazard curve at specific IM levels, multiplying by the increment in IMIM), and summing the results across all IM levels. This procedure is expressed in Equation (12) and visually depicted in Figure 15.
λ c = i = 1 P C | i m i · d λ I M i m i d i m · Δ i m

6.1. Relationship between λc and Probability of Collapse

λc denotes the mean annual frequency of furniture overturning. Assuming that earthquake occurrences follow a Poisson process over time, the probability of at least one overturning event within a period of t years can be determined using the following equation:
P c i n   t   y e a r s = 1 e x p λ c t
Given that λc is typically a small value for most buildings, the annual probability of overturning is approximately equal to λc, expressed as follows:
P c i n   1   y e a r s λ c

6.2. Deaggregation of λc

λc alone does not reveal which ground motion intensities contribute most to the collapse risk. To address this, a deaggregation of λc provides a valuable method for identifying the relative contributions of various ground motion intensities to the overall collapse risk. This process corresponds to the deaggregation by magnitude, distance, and epsilon (ε) commonly used in PSHA to determine the primary sources contributing to the hazard. The parameter ε quantifies the deviation, in terms of logarithmic standard deviations, between the observed ground spectral acceleration and the spectral acceleration predicted by an attenuation relationship at an arbitrary period.
A point on the λc deaggregation curve is retrieved as the product of the overturning probability at a ground motion intensity and the slope of the hazard curve as an intensity function. As indicated in Equation (12), the contribution of a specific (narrow) range of ground motion intensities to λc is calculated by multiplying the collapse probability at the midpoint intensity of the range by the slope of the hazard curve at the intensity and by the width of the intensity, ΔIM. As illustrated in Figure 15, the total area under the deaggregation curve equals λc. Ground motion intensities with higher values on the deaggregation curve represent more remarkable contributions to λc.
Figure 15 demonstrates that ground motion intensities associated with high overturning probabilities do not always contribute significantly to λc, as these intensities occur less frequently than those in the lower half of the overturning fragility curve. Consequently, the most significant contribution to λc generally comes from intensities within the lower half of the fragility curve, where the collapse probabilities may be smaller. Still, the seismic hazard curve’s slope is typically steeper compared to that for higher intensities.

6.3. Hazard Curve Deaggregation of λc

As one of the most earthquake-prone regions on the planet, Japan is chosen as the case study. The PSHA requires a hazard curve in spectral form. The National Research Institute for Earth Science and Disaster Resilience (NIED) provides the Japan Seismic Hazard Information (J-SHIS) system [147,148]. J-SHIS has developed a database of hazard curves based on the response spectrum for exceedance probabilities corresponding to a 50-year return period. These hazard curves are constructed by combining various scenarios involving subduction zones and near-fault earthquakes.
The damping factor provided by J-SHIS is set at 5%, which is generally applicable to reinforced concrete (RC) structures. However, the typical damping factor for steel structures is 2%, as used in the simulations conducted in this research. Therefore, the spectral response data from J-SHIS must be adjusted to reflect the corresponding damping factor. For this conversion, the following equation is provided by AIJ [5].
D h = 1 + 25 h 0 1 + 25 h e q
where h0 is the initial damping factor, and heq is the equivalent damping factor. Additionally, the probability for a return period of Tc is converted to the mean annual frequency using the following equation.
λ c = l n 1 P c T c
where Pc is the probability of exceedance and Tc designates the time interval where the probability Pc is calculated [149].
Additionally, the hazard curve is provided for specific natural periods: 0.1 s, 0.2 s, 0.3 s, 0.5 s, 1.0 s, 2.0 s, 3.0 s, and 5.0 s. To derive the hazard curve corresponding to the natural periods of the MRF and BRBF systems, the hazard curves are linearly interpolated in log-log space. Since the hazard curves in J-SHIS are derived from empirical equations using arbitrary constants, theoretical interpolation is not possible. Instead, previous studies have customarily interpolated the hazard curve in log-log space [69,150]. Therefore, the interpolation method used in this study follows earlier investigations [69,150].
The target locations are listed in Table 4, which includes sites selected from various regions across Japan. The hazard curves retrieved from J-SHIS are summarized in Figure 16. The frequency generally becomes greater in a shorter period. The discrepancy varies depending on the region, and Aichi exhibits a relatively huge gap among the natural period.
Figure 17a illustrates the mean annual frequency for low-height furniture. Overall, the MRF demonstrates the highest mean annual frequency except in Aichi. The greatest frequency is observed in the case of θt = 1/200 rad (γIag = 3.6 m/s2). The hazard curve in Figure 16 shows that Aichi has a higher probability of occurrence in a shorter natural period. Due to the seismic retrofitting principle using BRB, the natural period shortens. As Aichi’s seismic motion characteristics cause more significant acceleration responses in buildings with shorter natural periods, the computed mean annual frequency can exceed that of buildings without seismic retrofitting.
Additionally, the risk of furniture overturning at θt = 1/200 rad (γIag = 3.0 m/s2) and θt = 1/200 rad (γIag = 4.2 m/s2) is nearly identical in most cases. As shown in Figure 11, the acceleration response increases in θt = 1/200 rad (γIag = 4.2 m/s2) due to greater stiffness, while the threshold acceleration also increases. Consequently, the probability of furniture overturning is lowest in θt = 1/200 rad (γIag = 4.2 m/s2) according to the fragility curve in Figure 14. Meanwhile, the probability of exceedance in acceleration response tends to be more pronounced at shorter natural periods. As previously mentioned, the mean annual frequency is calculated as the product of the fragility curve and the derivative of the hazard curve. Thus, the effectiveness of seismic retrofitting with BRB depends on the balance between the reduction in fragility and the resonance with the seismic wave.
Figure 17b presents the case for medium-height furniture, where the risk of overturning is nearly the same for MRF and θt = 1/120 rad. According to the fragility curve in Figure 17b, fragility is reduced through seismic retrofitting. However, the spectral response increases due to the shorter natural period, resulting in nearly identical computed mean annual frequencies. The lowest frequency is observed in θt = 1/200 rad (γIag = 4.2 m/s2) across all locations, with the second lowest in θt = 1/200 rad (γIag = 3.0 m/s2). The second-most stringent design criterion of θt = 1/200 rad (γIag = 3.6 m/s2) exceed the aforementioned two cases and is highest in Aichi.
Figure 17c displays the mean annual frequency for tall furniture. The difference between retrofitting cases is less pronounced compared to low and medium-height furniture. The magnitude relation remains consistent with Figure 17b. However, the lowest mean annual frequency is achieved by MRF, owing to a more significant hazard curve at shorter natural periods. The mean annual frequency reaches 3.8 times greater in θt = 1/200 rad (γIag = 3.6 m/s2) compared to the MRF.

6.4. Discussion

Seismic retrofitting with BRBs effectively reduces displacement response and prevents building collapse. The design procedure has become straightforward due to the use of equivalent linearization. This method begins with determining the target story drift, making displacement response the primary concern. However, the risk of furniture overturning can increase as the building’s natural period shortens due to the added stiffness from the BRBs. In particular, the mean annual frequency increases when the spectral response at shorter periods becomes more significant. The proposed method quantifies this risk, enabling engineers to consider BCP when designing BRBFs for seismic retrofitting.
This study assumes that the furniture is not anchored. The fragility of furniture overturning can be mitigated through proper anchorage. Even with the installation of seismic damping devices, careful attention is required to maintain business continuity.

7. Conclusions

This study applies probabilistic risk assessment (PRA) to the issue of furniture overturning. Buckling-restrained braced frames (BRBFs) are designed using various design criteria based on the moment-resisting frame (MRF) provisions. Incremental dynamic analysis (IDA) is performed using artificial seismic waves that reflect realistic ground conditions. The resulting fragility curve for furniture overturning is then compared with hazard curves across Japan. The findings are summarized as follows:
(1)
Peak story drift is effectively mitigated with more stringent BRB design criteria. However, when the target story drift is reduced, peak floor acceleration can increase in BRBFs compared to MRFs.
(2)
The fragility of furniture overturning decreases with more stringent design criteria due to an increased critical acceleration threshold.
(3)
The mean annual frequency of acceleration response spectra generally increases with shorter natural periods. Consequently, the mean annual frequency of furniture overturning can be higher in BRBFs compared to MRFs.
(4)
The risk of furniture overturning is influenced by the balance between the building’s performance and the specific hazard characteristics at the location. Particular attention should be given to the shape of the hazard curve, especially in cases where short-period vibrations are prominent.
(5)
The calculation flow outlined in this research provides engineers and researchers with a clear methodology to quantify the risk of furniture overturning in a precise and explicit manner.
The model structure used in this research is a four-story building. In future research, it will be essential to assess the potential risk of furniture overturning in high-rise structures to extend safety warnings to the broader community. Future studies will explore the application of this methodology to buildings with varying aspect ratios and address additional business continuity planning (BCP) concerns, such as roof collapses. A more comprehensive investigation will be presented in subsequent papers.

Author Contributions

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

Funding

This research was funded by JSPS KAKENHI (Grant Number 24H00342) (Principal Investigator: Dr. Yoshihiro Kimura), JSPS KAKENHI (Grant Number 23K13392) (Principal Investigator: Dr. Atsushi Suzuki), and The Japan Iron and Steel Federation (Principal Investigator: Dr. Atsushi Suzuki). We express our deepest gratitude for their sincere support.

Data Availability Statement

The raw/processed data necessary to reproduce these findings cannot be shared at this time because the data also form part of an ongoing study.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Kimura, K.; Yoshioka, K.; Takeda, T.; Furuya, N.; Takemoto, Y. An experimental study of braces encased in steel tube and mortar. In Proceedings of the Architectural Institute of Japan Annual Meeting, Hokkaido, Japan, 15–17 September 1976; pp. 1041–1042. [Google Scholar]
  2. Takeuchi, T.; Hajjar, J.F.; Matsui, R.; Nishimoto, K.; Aiken, I.D. Local buckling restraint condition for core plates in buckling restrained braces. J. Constr. Steel Res. 2010, 66, 139–149. [Google Scholar] [CrossRef]
  3. Takeuchi, T.; Ida, M.; Yamada, S.; Suzuki, K. Estimation of cumulative deformation capacity of buckling restrained braces. J. Struct. Eng. 2008, 134, 822–831. [Google Scholar] [CrossRef]
  4. Takeuchi, T.; Ozaki, H.; Matsui, R.; Sutoh, F. Out-of-plane stability of buckling-restrained braces including moment transfer capacity. Earthq. Eng. Struct. Dyn. 2014, 43, 851–869. [Google Scholar] [CrossRef]
  5. Architectural Instutite of Japan (AIJ). Recommended Provisions for Seismic Damping Systems Applied to Steel Structures; Maruzen Publishing Co., Ltd.: Tokyo, Japan, 2014. (In Japanese) [Google Scholar]
  6. American Instutite of Steel Construction (AISC). Seismic Provisions for Structural Steel Buildings; American Instutite of Steel Construction (AISC): Chicago, IL, USA, 2016. [Google Scholar]
  7. Kasai, K.; Fu, Y.; Watanabe, A. Passive control systems for seismic damage mitigation. J. Struct. Eng. 1998, 124, 501–512. [Google Scholar] [CrossRef]
  8. Kimura, Y.; Suzuki, A.; Kasai, K. Estimation of plastic deformation capacity for H-shaped beams on local buckling under compressive and tensile forces. J. Struct. Constr. Eng. (Trans. AIJ) 2016, 81, 2133–2142. (In Japanese) [Google Scholar] [CrossRef]
  9. Suzuki, A.; Kimura, Y.; Kasai, K. Plastic deformation capacity of h-shaped beams collapse with combined buckling under reversed axial forces. J. Struct. Constr. Eng. (Trans. AIJ) 2018, 83, 297–307. (In Japanese) [Google Scholar] [CrossRef]
  10. Suzuki, A.; Kimura, Y.; Kasai, K. Rotation capacity of I-shaped beams under alternating axial forces based on buckling-mode transitions. J. Struct. Eng. 2020, 146, 04020089. [Google Scholar] [CrossRef]
  11. Kimura, Y.; Fujak, M.S.; Suzuki, A. Elastic local buckling strength of I-beam cantilevers subjected to bending moment and shear force based on flange–web interaction. Thin-Walled Struct. 2021, 162, 107633. [Google Scholar] [CrossRef]
  12. Kimura, Y.; Sato, Y.; Suzuki, A. Effect of fork restraint of column on lateral buckling behavior for H-shaped steel beams with continuous braces under flexural moment gradient. J. Struct. Constr. Eng. (Trans. AIJ) 2022, 87, 316–327. (In Japanese) [Google Scholar] [CrossRef]
  13. Fujak, M.S.; Kimura, Y.; Suzuki, A. Estimation of elastoplastic local buckling capacities and novel classification of I-beams based on beam’s elastic local buckling strength. Structures 2022, 39, 765–781. [Google Scholar] [CrossRef]
  14. Suzuki, A.; Kimura, Y. Rotation capacity of I-shaped beam failed by local buckling in buckling-restrained braced frames with rigid beam-column connections. J. Struct. Eng. 2023, 149, 04022243. [Google Scholar] [CrossRef]
  15. Fujak, M.S.; Suzuki, A.; Kimura, Y. Estimation of ultimate capacities of moment-resisting frame’s subassemblies with mid-story pin connection based on elastoplastic local buckling. Structures 2023, 48, 410–426. [Google Scholar] [CrossRef]
  16. Suzuki, A.; Kimura, Y.; Matsuda, Y.; Kasai, K. Rotation capacity of I-shaped beams with concrete slab in buckling-restrained braced frames. J. Struct. Eng. 2024, 150, 04023204. [Google Scholar] [CrossRef]
  17. Suzuki, A.; Kimura, Y. Rotation capacity of I-beams under cyclic loading with different kinematic/isotropic hardening characteristics. J. Constr. Steel Res. 2024, 223, 109007. [Google Scholar] [CrossRef]
  18. Aiken, I.D.; Mahin, S.A.; Uriz, P. Large-scale testing of buckling restrained braced frames. In Proceedings of the Japan Passive Control Symposium; Tokyo Institute of Technology: Tokyo, Japan, 2002. [Google Scholar]
  19. Berman, J.W.; Bruneau, M. Cyclic testing of a buckling restrained braced frame with unconstrained gusset connections. J. Struct. Eng. 2009, 135, 1499–1510. [Google Scholar] [CrossRef]
  20. Fahnestock, L.A.; Sause, R.; Ricles, J.M. Seismic response anad performance of buckling-restrained braced frames. J. Struct. Eng. 2007, 133, 1195–1204. [Google Scholar] [CrossRef]
  21. Hsiao, P.C.; Lehman, D.E.; Roeder, C.W. Improved analytical model for special concentrically braced frames. J. Constr. Steel Res. 2012, 73, 80–94. [Google Scholar] [CrossRef]
  22. Jia, M.; Lu, D.; Guo, L.; Sub, L. Experimental research and cyclic behavior of buckling-restrained braced composite frame. J. Constr. Steel Res. 2014, 95, 90–105. [Google Scholar] [CrossRef]
  23. Lin, P.C.; Tsai, K.C.; Wang, K.J.; Yu, Y.J.; Wei, C.Y.; Wu, A.C.; Tsai, C.Y.; Lin, C.H.; Chen, J.C.; Schellenberg, A.H.; et al. Seismic design and hybrid tests of a full-scale three-story buckling-restrained braced frame using welded end connections and thin profile. Earthq. Eng. Struct. Dyn. 2012, 41, 1001–1020. [Google Scholar] [CrossRef]
  24. Mcmanus, P.S.; Maxmahon, A.; Puckett, J. Buckling restrained braced frame with all-bolted gusset connections. Eng. J. 2013, 50, 89–116. [Google Scholar] [CrossRef]
  25. Palmer, K.D.; Christopulos, A.S.; Lehman, D.E.; Roeder, C.W. Experimental evaluation of cyclically loaded, large-scale, planar and 3-D buckling restrained braced frames. J. Constr. Steel Res. 2014, 101, 415–425. [Google Scholar] [CrossRef]
  26. Qu, Z.; Xie, J.; Cao, Y.; Li, W.; Wang, T. Effect of strain rate on the hysteretic behavior of buckling-restrained braces. J. Struct. Eng. 2020, 146, 06019003. [Google Scholar] [CrossRef]
  27. Tsai, K.; Hsiao, K.P.; Wang, K.; Weng, Y.; Lin, M.; Lin, K.; Chen, C.; Lai, J.; Lin, S. Pseudo-dynamic tests of a full-scale CFT/BRB frame—Part I: Specimen design, experiment and analysis. Earthq. Eng. Struct. Dyn. 2008, 37, 1081–1098. [Google Scholar] [CrossRef]
  28. Tsai, K.; Hsiao, P. Pseudo-dynamic test of a full scale CFT/BRB frame–Part II: Seismic performance of buckling-restrained braces and connections. Earthq. Eng. Struct. Dyn. 2008, 37, 1099–1115. [Google Scholar] [CrossRef]
  29. Ozkilic, Y.O.; Bozkurt, M.B.; Topkaya, C. Evaluation of seismic response factors for BRBFs using FEMA P695 methodology. J. Constr. Steel Res. 2018, 151, 41–57. [Google Scholar] [CrossRef]
  30. Ozkilic, 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]
  31. Ozkilic, Y.O.; Bozkurt, M.B. Numerical validation on novel replaceable reduced beam section connections for moment-resisting frames. Structures 2023, 50, 63–79. [Google Scholar] [CrossRef]
  32. Lignos, D.G.; Krawinkler, H. Development and utilization of structural component databases for performance-based earthquake engineering. J. Struct. Eng. 2013, 139, 1382–1394. [Google Scholar] [CrossRef]
  33. Chenouda, M.; Ayoub, A. Probabilistic collapse analysis of degrading Multi Degree of Freedom structures under earthquake excitation. Eng. Struct. 2009, 31, 2909–2921. [Google Scholar] [CrossRef]
  34. Scozzese, F.; Tubaldi, E.; Dall’Asta, A. Assessment of the effectiveness of multiple-stripe analysis by using a stochastic earthquake input model. Bull. Earthq. Eng. 2020, 18, 3167–3203. [Google Scholar] [CrossRef]
  35. Bradley, C.R.; Fahnestock, L.A.; Hine, E.M.; Sizemore, J.G. Full-scale cyclilc testing of low-ductility concentrically braced frames. J. Struct. Eng. 2017, 143, 04017029. [Google Scholar] [CrossRef]
  36. Chida, H.; Takahashi, N. Study on image diagnosis of timber houses damaged by earthquake using deep learning. Jpn. Archit. Rev. 2021, 4, 420–430. [Google Scholar] [CrossRef]
  37. Chang, P.C.; Flatau, A.; Liu, S.C. Review paper: Health monitoring of civil infrastructure. Struct. Health Monit. 2003, 2, 257–267. [Google Scholar] [CrossRef]
  38. Carden, E.P.; Fanning, P. Vibration based condition monitoring: A review. Struct. Health Monit. 2004, 3, 355–377. [Google Scholar] [CrossRef]
  39. Ji, X.; Fenves, G.L.; Kajiwara, K.; Nakashima, M. Seismic damage detection of a full-scale shaking table test structure. J. Struct. Eng. 2011, 137, 14–21. [Google Scholar] [CrossRef]
  40. Okada, K.; Shiroishi, R.; Morii, Y.; Sagawa, R. Method of structural health monitoring after earthquake using limited accelerometer—Case study of large-scale shaking table test on six story RC building. Concr. J. 2017, 55, 138–145. (In Japanese) [Google Scholar] [CrossRef]
  41. Gislason, G.P.; Mei, Q.; Gul, M. Rapid and automated damage detection in buildings through ARMAX analysis of wind induced vibrations. Front. Built Environ. 2019, 5, 16. [Google Scholar] [CrossRef]
  42. Alampalli, S.; Fu, G.; Dillon, E.W. Signal versus noise in damage detection by experimental modal analysis. J. Struct. Eng. 1997, 123, 237–245. [Google Scholar] [CrossRef]
  43. Ju, M.; Dou, Z.; Li, J.; Qiu, X.; Shen, B.; Zhang, S.; Yao, F.; Gong, W.; Wang, K. Piezoelectric materials and sensors for structural health monitoring: Fundamental aspects, current status, future perspectives. Sensors 2023, 23, 543. [Google Scholar] [CrossRef]
  44. Kishiki, S.; Iwasaki, Y. Evaluation of residual strength based on local buckling deformation of steel column—Quick inspection method for steel structures based on the visible damage Part 3. J. Struct. Constr. Eng. (Trans. AIJ) 2017, 82, 735–743. (In Japanese) [Google Scholar] [CrossRef]
  45. Gong, F.; Maekawa, K. Multi-scale simulation of freeze-thaw damage to RC column and its restoring force characteristics. Eng. Struct. 2018, 156, 522–536. [Google Scholar] [CrossRef]
  46. Gong, F.; Wang, Z.; Xia, J.; Maekawa, K. Coupled thermos-hydro-mechanical analysis of reinforced concrete beams under the effect of frost damage and sustained load. Struct. Concr. 2021, 22, 3430–3445. [Google Scholar] [CrossRef]
  47. Shimoi, N.; Nishida, T.; Obata, A.; Nakasho, K.; Cuadra, C. Comparison of displacement measurements in an exposed type column base using piezoelectric vibration sensors and piezoelectric limit sensors. Akita Prefect. Univ. Web J. B 2019, 6, 27–36. [Google Scholar]
  48. Aabid, A.; Parveez, B.; Raheman, M.A.; Ibrahim, Y.E.; Anjum, A.; Hrairi, M.; Parveen, N.; Zayan, J.M. A review of piezoelectric material-based structural control and health monitoring techniques for engineering structures: Challenges and opportunities. Actuators 2021, 10, 101. [Google Scholar] [CrossRef]
  49. Harada, T.; Yokoyama, K.; Tanabe, K. Study of crack detection of civil infrastructure using PVDF film sensor. J. Struct. Eng. 2013, 59A, 47–55. [Google Scholar]
  50. Shimoi, N.; Cuadra, C.; Madokoro, H.; Nakasho, K. Comparison of displacement measurements and simulation on fillet weld of steel column base. Int. J. Mech. Eng. Appl. 2020, 8, 111–117. [Google Scholar] [CrossRef]
  51. Suzuki, A.; Liao, W.; Shibata, D.; Yoshino, Y.; Kimura, Y.; Shimoi, N. Structural damage detection technique of secondary building components using piezoelectric sensors. Buildings 2023, 13, 2368. [Google Scholar] [CrossRef]
  52. Matsumoto, Y.; Yamada, S.; Iyama, J.; Koyama, T.; Kishiki, S.; Shimada, Y.; Asada, H.; Ikenaga, M. Damage to steel educational facilities in the 2011 East Japan Earthquake: Part 1 Outline of the reconnaissance and damage to major structural components. In Proceedings of the 15th World Conference on Earthquake Engineering, Lisbon, Portugal, 24–28 September 2012. [Google Scholar]
  53. Suzuki, A.; Fujita, T.; Kimura, Y. Identifying damage mechanisms of gymnasium structure damaged by the 2011 Tohoku earthquake based on biaxial excitation. Structures 2022, 35, 1321–1338. [Google Scholar] [CrossRef]
  54. Kaneko, M.; Hayashi, Y. Proposal of a curve to describe overturning ratios of rigid bodies. J. Struct. Constr. Eng. (Trans. AIJ) 2000, 536, 55–62. (In Japanese) [Google Scholar] [CrossRef]
  55. Kaneko, M. Method to estimate overturning ratios of furniture during earthquakes. J. Struct. Constr. Eng. (Trans. AIJ) 2002, 551, 61–68. (In Japanese) [Google Scholar] [CrossRef]
  56. Zhang, J.; Makris, N. Rocking response of free-standing blocks under cycloidal pulses. J. Eng. Mech. 2001, 127, 473–483. [Google Scholar] [CrossRef]
  57. Kounadis, A.N. Parametric study in rocking instability of a rigid block under harmonic ground pulse: A unified approach. Soil Dyn. Earthq. Eng. 2013, 45, 125–143. [Google Scholar] [CrossRef]
  58. Vassiliou, M.F.; Mackie, K.R.; Stojadinovic, B. Dynamic response analysis of solitary flexible rocking bodies: Modeling and behavior under pulse-like ground excitation. Earthq. Eng. Struct. Dyn. 2014, 43, 1463–1481. [Google Scholar] [CrossRef]
  59. Fragiadakis, M.; Diamantopoulos, S. Fragility and risk assessment of freestanding building contents. Earthq. Eng. Struct. Dyn. 2020, 49, 1028–1048. [Google Scholar] [CrossRef]
  60. Avgenakis, E.; Psycharis, I. An integrated macroelement formulation for the dynamic response of inelastic deformable rocking bodies. Earthq. Eng. Struct. Dyn. 2020, 49, 1072–1094. [Google Scholar] [CrossRef]
  61. Egidio, A.D.; Leo, A.M. Improvement of the dynamic and seismic response of non-structural rocking bodies through the ability to change their geometrical configuration. Eng. Struct. 2023, 275, 115231. [Google Scholar] [CrossRef]
  62. Saito, K.; Higuchi, T.; Nakazawa, S. Study of seismic risk evaluation consideration of overturning furniture for middle and low-rise buildings such as office buildings. Steel Constr. Eng. 2016, 23, 77–88. (In Japanese) [Google Scholar]
  63. Ishikawa, Y.; Takeda, M.; Okumura, T.; Hayashi, Y.; Kanegawa, S. Procedure for seismic risk evaluation of buildings. AIJ J. Technol. Des. 2000, 6, 275–278. (In Japanese) [Google Scholar] [CrossRef]
  64. Elkady, A.; Lignos, S.G. Effect of gravity framing on the overstrength and collapse capacity of steel frame buildings with perimeter special moment frames. Earthq. Eng. Struct. Dyn. 2015, 44, 1289–1307. [Google Scholar] [CrossRef]
  65. Jisr, H.E.; Lignos, D.G. Fragility assessment of beam-slab connections for informing earthquake-induced repairs in composite-steel moment resisting frames. Front. Built Environ. 2021, 7, 691553. [Google Scholar]
  66. Jisr, H.E.; Kohrangi, M.; Lignos, D.G. Proposed nonlinear macro-model for seismic risk assessment of composite-steel moment resisting frames. Earthq. Eng. Struct. Dyn. 2022, 51, 1180–1200. [Google Scholar] [CrossRef]
  67. Paranonessso, M.; Lignos, D.G. Seismic design and performance of steel concentrically braced frames buildings with dissipative floor connectors. Earthq. Eng. Struct. Dyn. 2022, 51, 3505–3525. [Google Scholar] [CrossRef]
  68. Vamvatsikos, D.; Cornell, C.A. Incremental dynamic analysis. Earthq. Eng. Struct. Dyn. 2002, 31, 491–514. [Google Scholar] [CrossRef]
  69. Eads, L.; Miranda, E.; Krawinkler, H.; Lignos, D. An efficient method for estimating the collapse risk of structures in seismic regions. Earthq. Eng. Struct. Dyn. 2013, 42, 25–41. [Google Scholar] [CrossRef]
  70. Deylami, A.; Mahdavipour, M.A. Probabilistic seismic demand assessment of residual drift for buckling-restrained braced frames as a dual system. Struct. Saf. 2016, 58, 31–39. [Google Scholar] [CrossRef]
  71. Architectural Instutite of Japan (AIJ). AIJ Standard for Allowable Stress Design of Steel Structures; Maruzen Publishing Co., Ltd.: Tokyo, Japan, 2019. (In Japanese) [Google Scholar]
  72. European Committee for Standardization (CEN). Design of Structures for Earthquake Resistance—Part I: General Rules, Seismic Actions and Rules for Buildings. Eurocode 8. 2003. CEN. Available online: https://www.confinedmasonry.org/wp-content/uploads/2009/09/Eurocode-8-1-Earthquakes-general.pdf (accessed on 4 October 2024). Eurocode 8, CEN..
  73. Architectural Instutite of Japan (AIJ). Recommendations for Loads on Buildings; Maruzen Publishing Co., Ltd.: Tokyo, Japan, 2015. (In Japanese) [Google Scholar]
  74. Dassault Systems. Simulia User Assistance 2024; Dassault Systems: Vélizy-Villacoublay, France, 2024. [Google Scholar]
  75. Architectural Instutite of Japan (AIJ). Design Recommendations for Composite Constructions; Maruzen Publishing Co., Ltd.: Tokyo, Japan, 2010. (In Japanese) [Google Scholar]
  76. Suzuki, A.; Kimura, Y. Cyclic behavior of component model of composite beam subjected to fully reversed cyclic loading. J. Struct. Eng. 2019, 145, 04019015. [Google Scholar] [CrossRef]
  77. Suzuki, A.; Abe, K.; Kimura, Y. Restraint performance of stud connection during lateral-torsional buckling under synchronized in-plane displacement and out-of-plane rotation. J. Struct. Eng. 2020, 146, 04020029. [Google Scholar] [CrossRef]
  78. Suzuki, A.; Abe, K.; Suzuki, K.; Kimura, Y. Cyclic behavior of perfobond shear connectors subjected to fully reversed cyclic loading. J. Struct. Eng. 2021, 147, 04020355. [Google Scholar] [CrossRef]
  79. Suzuki, A.; Suzuki, K.; Kimura, Y. Ultimate shear strength of perfobond shear connectors subjected to fully reversed cyclic loading. Eng. Struct. 2021, 248, 113240. [Google Scholar] [CrossRef]
  80. Suzuki, A.; Kimura, Y. Mechanical performance of stud connection in steel–concrete composite beam under reversed stress. Eng. Struct. 2021, 249, 113338. [Google Scholar] [CrossRef]
  81. Suzuki, A.; Hiraga, K.; Kimura, Y. Cyclic behavior of steel-concrete composite dowel by clothoid-shaped shear connectors under fully reversed cyclic stress. J. Adv. Concr. Technol. 2023, 21, 76–91. [Google Scholar] [CrossRef]
  82. Suzuki, A.; Hiraga, K.; Kimura, Y. Mechanical performance of puzzle-shaped shear connectors subjected to fully reversed cyclic stress. J. Struct. Eng. 2023, 149, 04023087. [Google Scholar] [CrossRef]
  83. Suzuki, A.; Hiraga, K.; Kimura, Y.; Takahashi, J.; Kikuchi, K. Evaluation of superiority mechanical shear connectors based on flexural stiffness of composite beam. AIJ J. Technol. Des. 2024, 30, 154–159. (In Japanese) [Google Scholar] [CrossRef]
  84. Ollgaard, J.G.; Slutter, R.G.; Fisher, J.W. Shear strength of stud connectors in lightweight and normal weight concrete. AISC Eng. J. 1971, 71, 55–64. [Google Scholar] [CrossRef]
  85. Hawkins, N.M.; Mitchell, D. Seismic response of composite shear connections. J. Struct. Eng. 1984, 110, 2120–2136. [Google Scholar] [CrossRef]
  86. Oehlers, D.J. Deterioration in strength of stud connectors in composite bridge beams. J. Struct. Eng. 1990, 116, 3417–3431. [Google Scholar] [CrossRef]
  87. Bursi, O.S.; Gramola, G. Behaviour of headed stud shear connectors under low-cycle high amplitude displacements. Mater. Struct. 1999, 32, 290–297. [Google Scholar] [CrossRef]
  88. Lin, W.; Yoda, T.; Taniguchi, N. Fatigue tests on straight steel-concrete composite beams subjected to hogging moment. J. Constr. Steel Res. 2013, 80, 42–56. [Google Scholar] [CrossRef]
  89. Lin, W.; Yoda, T.; Taniguchi, N.; Kasano, H.; He, J. Mechanical performance of steel-concrete composite beams subjected to a hogging moment. J. Struct. Eng. 2014, 140, 04013031. [Google Scholar] [CrossRef]
  90. Song, A.; Wan, S.; Jiang, Z.; Xu, J. Residual deflection analysis in negative moment regions of steel-concrete composite beams under fatigue loading. Constr. Build. Mater. 2018, 158, 50–60. [Google Scholar] [CrossRef]
  91. Japan Society of Civil Engineers (JSCE). Standard Specification for Hybrid Structures–2014; Maruzen Publishing Co., Ltd.: Tokyo, Japan, 2015. [Google Scholar]
  92. Oguejiofor, E.C.; Hosain, M.U. Behavior of perfobond rib shear connectors in composite beams. Can. J. Civ. Eng. 1992, 19, 1–10. [Google Scholar] [CrossRef]
  93. Oguejiofor, E.C.; Hosain, M.U. Numerical analysis of pushout specimens with perfobond rib connectors. Comput. Struct. 1997, 62, 617–624. [Google Scholar] [CrossRef]
  94. Kim, S.H.; Kim, K.S.; Park, S.; Ahn, J.H.; Lee, M.K. Y-type perfobond rib shear connectors subjected to fatigue loading on highway bridges. J. Constr. Steel Res. 2016, 122, 445–454. [Google Scholar] [CrossRef]
  95. Kim, S.H.; Kim, K.S.; Han, O.; Park, J.S. Influence of transverse rebar on shear behavior of Y-type perfobond rib shear connection. Constr. Build. Mater. 2018, 180, 254–264. [Google Scholar] [CrossRef]
  96. Kim, S.H.; Batbold, T.; Shah, S.H.A.; Yoon, S.; Han, O. Development of shear resistance formula for the Y-type perfobond rib shear connector considering probabilistic characteristics. Appl. Sci. 2021, 11, 3877. [Google Scholar] [CrossRef]
  97. Tanaka, T.; Sakai, J.; Kawano, A. Development of new mechanical shear connector using burring steel plate. J. Struct. Constr. Eng. (Trans. AIJ) 2013, 78, 2237–2245. [Google Scholar] [CrossRef]
  98. Tanaka, T.; Sakai, J.; Kawano, A. Experimental study on elastic-plastic flexural behavior of composite beam by burring shear connector and perfobond-rib shear connector. Steel Constr. Eng. 2014, 21, 111–123. [Google Scholar]
  99. Tanaka, T.; Norimatsu, K.; Sakai, J.; Kawano, A. Development of innovative shear connector in steel-concrete joint and establishment of rational design method. Proc. Jpn. Concr. Inst. 2013, 35, 1237–1242. [Google Scholar]
  100. Lorenc, W.; Kozuch, M.; Rowinski, S. The behaviour of puzzle-shaped composite dowels—Part I: Experimental study. J. Constr. Steel Res. 2014, 101, 482–499. [Google Scholar] [CrossRef]
  101. Lorenc, W.; Kozuch, M.; Rowinski, S. The behaviour of puzzle-shaped composite dowels—Part II: Theoretical investigations. J. Constr. Steel Res. 2014, 101, 500–518. [Google Scholar] [CrossRef]
  102. Kopp, M.; Wolters, K.; Classen, M.; Hegger, J.; Gundel, M.; Gallwoszus, J.; Heinemeyer, S.; Feldmann, M. Composite dowels as shear connectors for composite beams—Background to the design concept for static loading. J. Constr. Steel Res. 2018, 147, 488–503. [Google Scholar] [CrossRef]
  103. Lorenc, W. Concrete failure of composite dowels under cyclic loading during full-scale tests of beams for the “Wierna Rzeka” bridge. Eng. Struct. 2020, 209, 110199. [Google Scholar] [CrossRef]
  104. Classen, M.; Hegger, J. Shear-slip behaviour and ductility of composite dowel connectors with pry-out failure. Eng. Struct. 2017, 150, 428–437. [Google Scholar] [CrossRef]
  105. Lorenc, W.; Kurz, W.; Seidl, G. Hybrid steel-concrete sections for bridges: Definition and basis for design. Eng. Struct. 2022, 270, 114902. [Google Scholar] [CrossRef]
  106. Christou, G.; Hegger, J.; Classen, M. Fatigue of clothoid shaped rib shear connectors. J. Constr. Steel Res. 2020, 171, 106133. [Google Scholar] [CrossRef]
  107. Classen, M.; Gallwoszus, J. Concrete fatigue in composite dowels. Struct. Concr. 2016, 17, 63–73. [Google Scholar] [CrossRef]
  108. Classen, M.; Gallwoszus, J.; Stark, A. Anchorage of composite dowels in UHPC under fatigue loading. Struct. Concr. 2016, 17, 183–193. [Google Scholar] [CrossRef]
  109. Kim, H.Y.; Jeong, Y.J. Experimental investigation on behaviour of steel-concrete composite bridge decks with perfobond ribs. J. Constr. Steel Res. 2006, 62, 463–471. [Google Scholar] [CrossRef]
  110. Christou, G.; Classen, M.; Wolters, K.; Broschart, Y. Fatigue of composite constructions with composite dowels. In Proceedings of the 14th Nordic Steel Construction Conference, Copenhagen, Denmark, 18–20 September 2019; pp. 277–282. [Google Scholar]
  111. Kopp, M.; Christou, G.; Stark, A.; Hegger, J.; Feldmann, M. Integrated slab system for the steel and composite construction. Bauingenieur 2018, 87, 136–148. [Google Scholar]
  112. Japan Society of Steel Construction (JSSC). Guideline of Standard Push-Out Tests of Headed Stud and Current Situation of Research on Stud Shear Connectors; Japan Society of Steel Construction (JSSC): Tokyo, Japan, 1996. [Google Scholar]
  113. Ahmed, S.E. Ultimate Shear Resistance of Plate Girders Part 2—Hoglund Theory. Int. J. Civ. Environ. Eng. 2013, 7, 918–926. [Google Scholar]
  114. Alinia, M.M.; Shakiba, M.; Habashi, H.R. Shear Failure Characteristics of Steel Plate Girders. Thin-Walled Struct. 2009, 47, 1498–1506. [Google Scholar] [CrossRef]
  115. Anthimos, A.; Marius, M.; Victor, G. Prediction of available rotation capacity and ductility of wide-flange beams part 2 applications. J. Constr. Steel Res. 2012, 68, 176–191. [Google Scholar]
  116. Basler, K. Strength of plate girders in shear. J. Struct. Div. 1961, 87, 151–180. [Google Scholar] [CrossRef]
  117. Cheng, X.; Chen, Y.; Pan, L. Experimental study of steel beam-columns composed of slender H-sections under cyclic bending. J. Constr. Steel Res. 2013, 88, 279–288. [Google Scholar] [CrossRef]
  118. Elkady, A.; Lignos, D.G. Full-scale testing of deep wide-flange steel columns under multiaxis cyclic loading. J. Struct. Eng. 2018, 144, 2018. [Google Scholar] [CrossRef]
  119. Gioncu, V.; Marius, M.; Anthimos, A. Prediction of available rotation capacity and ductility of wide-flange beams part 1 DUCTROT-M computer program. J. Constr. Steel Res. 2012, 69, 8–19. [Google Scholar] [CrossRef]
  120. Glassman, D.J.; Garlock, M.E. A compression model for ultimate postbuckling shear strength. Thin-Walled Struct. 2016, 102, 258–272. [Google Scholar] [CrossRef]
  121. Gulen, O.; Harris, J.; Uang, C. Observations from cyclic tests on deep, wide-flange beam-columns. Eng. J. 2017, 54, 45–59. [Google Scholar]
  122. Hanna, M.T. Failure loads of web panels loaded in pure shear. J. Constr. Steel Res. 2015, 105, 39–48. [Google Scholar] [CrossRef]
  123. Ikarashi, K.; Suekuni, R.; Shinohara, T.; Wang, T. Evaluation of plastic deformation capacity of H-shaped steel beams with newly proposed limitation value of plate slenderness. J. Struct. Constr. Eng. (Trans. AIJ) 2011, 76, 1865–1872. (In Japanese) [Google Scholar] [CrossRef]
  124. Kasai, K.; Popov, E.P. General behavior of WF steel shear link beams. J. Struct. Eng. 1986, 112, 362–382. [Google Scholar] [CrossRef]
  125. Kasai, K.; Popov, E.P. Cyclic Web Buckling Control for Shear Link Beams. J. Struct. Eng. 1986, 112, 505–523. [Google Scholar] [CrossRef]
  126. Kato, B. Rotation capacity of H-section members as determined by local buckling. J. Constr. Steel Res. 1989, 13, 95–109. [Google Scholar] [CrossRef]
  127. Lee, C.S.; Davidson, J.S.; Yoo, C.H. Shear buckling coefficients of plate girder web panels. Comput. Struct. 1996, 59, 789–795. [Google Scholar] [CrossRef]
  128. Lee, C.D.; Lee, D.; Yoo, C.H. Flexure and Shear Interaction in Steel I-girders. J. Struct. Eng. 2013, 139, 1882–1894. [Google Scholar] [CrossRef]
  129. MacRae, G.A. The Seismic Response of Steel Frames. Ph.D. Thesis, University of Canterbury, Christchurch, New Zealand, 1999. [Google Scholar]
  130. Miguel, A.; Luis, M.; Jose, M.C. Evaluation of the rotation capacity limits of steel members defined in EC8-3. J. Constr. Steel Res. 2017, 135, 11–29. [Google Scholar]
  131. Nakashima, M.; Takanashi, K.; Kato, H. Test of steel beam-columns subject to sidesway. J. Struct. Eng. 1990, 116, 2516–2531. [Google Scholar] [CrossRef]
  132. Nowell, J.D.; Uang, C. Cyclic behavior of steel wide-flange columns subjected to large drift. J. Struct. Eng. 2008, 134, 1334–1342. [Google Scholar] [CrossRef]
  133. Shahabedding, T.; Schafer, B.W. Role of local slenderness in the rotation capacity of structural steel members. J. Constr. Steel Res. 2014, 95, 32–43. [Google Scholar]
  134. Shokouhian, M.; Shi, Y. Classification of I-section flexural members based on member ductility. J. Constr. Steel Res. 2014, 95, 198–210. [Google Scholar] [CrossRef]
  135. Suzuki, T.; Ikarashi, K.; Tsuneki, Y. A study of collapse mode and plastic deformation capacity of H-shaped steel beams under shear bending. J. Struct. Constr. Eng. (Trans. AIJ) 2001, 547, 185–191. (In Japanese) [Google Scholar] [CrossRef]
  136. Torsten, H. Simply supported long thin plate I-girders without web stiffeners subjected to distributed transverse load. IABSE Rep. Work. Comm. 1971, 85–97. [Google Scholar]
  137. Torsten, H. Shear buckling resistance of steel and aluminium plate girders. Thin-Walled Struct. 1997, 29, 13–30. [Google Scholar]
  138. Wael, F.R. Local buckling of welded steel I-beams considering flange–web interaction. Thin-Walled Struct. 2015, 97, 241–249. [Google Scholar]
  139. Yiyi, C.; Xin, C.; Nethercot, N.D.A. An overview study on cross-section classification of steel H-sections. J. Constr. Steel Res. 2013, 80, 386–393. [Google Scholar]
  140. The Headquarters for Earthquake Research Promotion. National Seismic Hazard Map for Japan. Available online: https://www.jishin.go.jp/evaluation/seismic_hazard_map/ (accessed on 5 September 2024).
  141. Japan Seismic Hazard Information Station (J-SHIS). Available online: https://www.j-shis.bosai.go.jp/ (accessed on 5 September 2024).
  142. Kamae, K.; Irikura, K. Prediction of site-specific strong ground motion using semi-empirical methods. In Proceeding of the 10th World Conference on Earthquake Engineering, Madrid, Spain, 19–24 July 1992; pp. 801–806. [Google Scholar]
  143. Boore, D.M. Stochastic simulation of high-frequency ground motions based on seismological models of the radiated spectra. Bull. Seismol. Soc. Am. 1983, 73, 1865–1984. [Google Scholar]
  144. Satoh, T.; Kawase, H.; Sato, T. Statistical spectral model of earthquakes in the eastern Tohoku district, Japan, based on the surface and borehole records observed in Sendai. Bull. Seismol. Soc. Am. 1997, 87, 446–462. [Google Scholar] [CrossRef]
  145. Sendai City. A distribution map of seismic intensity in JMA in the expected off Miyagi earthquake. Investig. Rep. Earthq. Damage Estim. 2002. [Google Scholar]
  146. Ohsaki, Y.; Hara, A.; Kiyota, Y. Stress-strain model for soils for seismic analysis. In Proceedings of the 5th Japan Earthquake Engineering Symposium, Tokyo, Japan, 28–30 November 1978; pp. 697–704. [Google Scholar]
  147. Fujiwara, H.; Morikawa, N.; Maeda, T.; Iwaki, A.; Senna, S.; Kawai, S.; Azuma, H.; Xiansheng, K.H.; Imoto, M.; Wakamatsu, K.; et al. Improved Hazard Assessment after the 2011 Great East Japan Earthquake (Part 2); National Research Institute for Earth Science and Disaster Resilience: Tsukuba, Japan, 2023. (In Japanese) [Google Scholar]
  148. Subcommittee on Strong Motion Evaluation of the Earthquake Research Committee. Seismic Hazard of Response Spectra (Prototype Version); National Research Institute for Earth Science and Disaster Resilience: Tsukuba, Japan. Available online: https://www.jishin.go.jp/evaluation/seismic_hazard_map/sh_response_spectrum/ (accessed on 4 October 2024). (In Japanese).
  149. Bantilas, K.E.; Naoum, M.C.; Kavvadias, I.E.; Karayannis, C.G.; Elenas, A. Structural pounding effect on the seismic performance of a multistorey reinforced concrete frame structure. Infrastructures 2023, 8, 122. [Google Scholar] [CrossRef]
  150. Garcia, J.R.; Miranda, E. Performance-Based Assessment of Existing Structures Accounting for Residual Displacements; Stanford University: Stanford, CA, USA, 2005. [Google Scholar]
Figure 1. Configuration of Buckling-Restrained Braced Frame (BRBF): (a) diagonal configuration; (b) V configuration.
Figure 1. Configuration of Buckling-Restrained Braced Frame (BRBF): (a) diagonal configuration; (b) V configuration.
Buildings 14 03195 g001
Figure 2. Concept of this research: (a) MRF; (b) BRBF; (c) furniture overturning after the 2022 Fukushima-Oki earthquake.
Figure 2. Concept of this research: (a) MRF; (b) BRBF; (c) furniture overturning after the 2022 Fukushima-Oki earthquake.
Buildings 14 03195 g002
Figure 3. Drawing and member schedule of model structure.
Figure 3. Drawing and member schedule of model structure.
Buildings 14 03195 g003
Figure 4. Idealized response spectra: (a) displacement; (b) velocity; (c) acceleration.
Figure 4. Idealized response spectra: (a) displacement; (b) velocity; (c) acceleration.
Buildings 14 03195 g004
Figure 5. Calibration of BRB specification: (a) displacement response spectrum; (b) performance curve.
Figure 5. Calibration of BRB specification: (a) displacement response spectrum; (b) performance curve.
Buildings 14 03195 g005
Figure 6. FEA model of model structure.
Figure 6. FEA model of model structure.
Buildings 14 03195 g006
Figure 7. Fragility curve of furniture overturning.
Figure 7. Fragility curve of furniture overturning.
Buildings 14 03195 g007
Figure 8. Earthquake scenarios concerned.
Figure 8. Earthquake scenarios concerned.
Buildings 14 03195 g008
Figure 9. Acceleration response spectra (h = 0.02).
Figure 9. Acceleration response spectra (h = 0.02).
Buildings 14 03195 g009
Figure 10. Maximum inter-story drift obtained from IDA: (a) MRF; (b) θt = 1/120 rad; (c) θt = 1/150 rad; (d) θt = 1/200 rad (γIag = 3.0 m/s2); (e) θt = 1/200 rad (γIag = 3.6 m/s2); (f) θt = 1/200 rad (γIag = 4.2 m/s2).
Figure 10. Maximum inter-story drift obtained from IDA: (a) MRF; (b) θt = 1/120 rad; (c) θt = 1/150 rad; (d) θt = 1/200 rad (γIag = 3.0 m/s2); (e) θt = 1/200 rad (γIag = 3.6 m/s2); (f) θt = 1/200 rad (γIag = 4.2 m/s2).
Buildings 14 03195 g010
Figure 11. Peak floor acceleration obtained from IDA: (a) MRF; (b) θt = 1/120 rad; (c) θt = 1/150 rad; (d) θt = 1/200 rad (γIag = 3.0 m/s2); (e) θt = 1/200 rad (γIag = 3.6 m/s2); (f) θt = 1/200 rad (γIag = 4.2 m/s2).
Figure 11. Peak floor acceleration obtained from IDA: (a) MRF; (b) θt = 1/120 rad; (c) θt = 1/150 rad; (d) θt = 1/200 rad (γIag = 3.0 m/s2); (e) θt = 1/200 rad (γIag = 3.6 m/s2); (f) θt = 1/200 rad (γIag = 4.2 m/s2).
Buildings 14 03195 g011
Figure 12. Threshold acceleration causing 50% of furniture overturning (low): (a) MRF; (b) θt = 1/120 rad; (c) θt = 1/150 rad; (d) θt = 1/200 rad (γIag = 3.0 m/s2); (e) θt = 1/200 rad (γIag = 3.6 m/s2); (f) θt = 1/200 rad (γIag = 4.2 m/s2).
Figure 12. Threshold acceleration causing 50% of furniture overturning (low): (a) MRF; (b) θt = 1/120 rad; (c) θt = 1/150 rad; (d) θt = 1/200 rad (γIag = 3.0 m/s2); (e) θt = 1/200 rad (γIag = 3.6 m/s2); (f) θt = 1/200 rad (γIag = 4.2 m/s2).
Buildings 14 03195 g012
Figure 13. Calculation flow of fragility curve of furniture overturning.
Figure 13. Calculation flow of fragility curve of furniture overturning.
Buildings 14 03195 g013
Figure 14. Fragility curve of furniture overturning: (a) low; (b) medium; (c) tall.
Figure 14. Fragility curve of furniture overturning: (a) low; (b) medium; (c) tall.
Buildings 14 03195 g014
Figure 15. Computation concept of mean annual frequency: (a) fragility curve; (b) derivative of hazard curve; (c) mean annual frequency.
Figure 15. Computation concept of mean annual frequency: (a) fragility curve; (b) derivative of hazard curve; (c) mean annual frequency.
Buildings 14 03195 g015
Figure 16. Hazard curves for locations across Japan: (a) Hokkaido; (b) Miyagi; (c) Tokyo; (d) Ishikawa; (e) Aichi; (f) Hyogo; (g) Hiroshima; (h) Kumamoto.
Figure 16. Hazard curves for locations across Japan: (a) Hokkaido; (b) Miyagi; (c) Tokyo; (d) Ishikawa; (e) Aichi; (f) Hyogo; (g) Hiroshima; (h) Kumamoto.
Buildings 14 03195 g016
Figure 17. Mean annual frequency of furniture overturning: (a) low; (b) medium; (c) tall.
Figure 17. Mean annual frequency of furniture overturning: (a) low; (b) medium; (c) tall.
Buildings 14 03195 g017
Table 1. Required capacities of BRBs (unit: kN/mm for Kfi and Kai, kN for Fai).
Table 1. Required capacities of BRBs (unit: kN/mm for Kfi and Kai, kN for Fai).
Θt1/120 rad1/150 rad1/200 rad1/200 rad1/200 rad
γIag3.0 m/s23.0 m/23.0 m/s23.6 m/s24.2 m/s2
StoryKfiKaiFaiKaiFaiKaiFaiKaiFaiKaiFai
R194.590.00450.2249.21246
4249.22.80018.60186.2936.9378.81906631.13175
3323.132.20216.1275.21385530.42669864.64350
2211.8198.52253270.22556475.23371690.54898972.46898
Table 2. Furniture profiles.
Table 2. Furniture profiles.
TypeB [mm]H [mm]Fb [Hz]Buildings 14 03195 i001
I (low)45.01200.883
II (medium)45.01500.859
III (tall)45.02000.814
Table 3. Profile of seismic waves (unit: m/s2 for PGA and m/s for PGV).
Table 3. Profile of seismic waves (unit: m/s2 for PGA and m/s for PGV).
SiteLongitudeLatitudeCaseDirectionPGAPGV
Aobayama38°15′18″ N140°50′18″ ECase 1NS7.770.64
EW6.550.58
Case2NS10.10.65
EW11.31.64
Katahira38°15′14″ N140°52′26″ ECase 1NS4.270.51
EW4.480.43
Case2NS6.230.48
EW6.421.21
Nagamachi38°13′26″ N140°52′49″ ECase 1NS5.630.54
EW5.250.52
Case2NS6.710.73
EW5.891.32
Table 4. Target locations for risk assessment.
Table 4. Target locations for risk assessment.
LocationJ-SHIS Mesh CodeLongitudeLatitudeElevation
Hokkaido644142781443.0615 N141.3547 E21 m
Miyagi574036292138.2677 N140.8703 E46 m
Tokyo5339452532 35.6885 N139.6922 E38 m
Ishikawa5436657223 36.5615 N136.6578 E27 m
Aichi5236671243 35.1823 N136.9078 E10 m
Hyogo5235012543 34.6906 N135.1953 E6 m
Hiroshima5132436612 34.3844 N132.4547 E2 m
Kumamoto4930156623 32.8031 N130.7078 E13 m
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

Suzuki, A.; Ohno, S.; Kimura, Y. Risk Assessment of Overturning of Freestanding Non-Structural Building Contents in Buckling-Restrained Braced Frames. Buildings 2024, 14, 3195. https://doi.org/10.3390/buildings14103195

AMA Style

Suzuki A, Ohno S, Kimura Y. Risk Assessment of Overturning of Freestanding Non-Structural Building Contents in Buckling-Restrained Braced Frames. Buildings. 2024; 14(10):3195. https://doi.org/10.3390/buildings14103195

Chicago/Turabian Style

Suzuki, Atsushi, Susumu Ohno, and Yoshihiro Kimura. 2024. "Risk Assessment of Overturning of Freestanding Non-Structural Building Contents in Buckling-Restrained Braced Frames" Buildings 14, no. 10: 3195. https://doi.org/10.3390/buildings14103195

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop