Next Article in Journal
Preventative Sensor-Based Remote Monitoring of the Diabetic Foot in Clinical Practice
Previous Article in Journal
Channel Prediction-Based Security Authentication for Artificial Intelligence of Things
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploration of Damage Identification Method for a Large-Span Timber Lattice Shell Structure in Taiyuan Botanical Garden based on Structural Health Monitoring

College of Civil Engineering, Taiyuan University of Technology, Taiyuan 030024, China
*
Author to whom correspondence should be addressed.
Sensors 2023, 23(15), 6710; https://doi.org/10.3390/s23156710
Submission received: 11 June 2023 / Revised: 21 July 2023 / Accepted: 24 July 2023 / Published: 27 July 2023
(This article belongs to the Topic Advances in Non-Destructive Testing Methods)

Abstract

:
Large-span spatial lattice structures generally have characteristics such as incomplete modal information, high modal density, and high degrees of freedom. To address the problem of misjudgment in the damage detection of large-span spatial structures caused by these characteristics, this paper proposed a damage identification method based on time series models. Firstly, the order of the autoregressive moving average (ARMA) model was selected based on the Akaike information criterion (AIC). Then, the long autoregressive method was used to estimate the parameters of the ARMA model and extract the residual sequence of the autocorrelation part of the model. Furthermore, principal component analysis (PCA) was introduced to reduce the dimensionality of the model while retaining the characteristic values. Finally, the Mahalanobis distance (MD) was used to construct the damage sensitive feature (DSF). The dome of Taiyuan Botanical Garden in China is one of the largest non-triangular timber lattice shells worldwide. Relying on the structural health monitoring (SHM) project of this structure, this paper verified the effectiveness of the damage identification model through numerical simulation and determined the damage degree of the dome structure through SHM measurement data. The results demonstrated that the proposed damage identification method can effectively identify the damage of large-span timber lattice structures, locate the damage position, and estimate the degree of damage. The constructed DSF had relatively strong robustness to small damage and environmental noise and has practical application value for SHM in engineering.

1. Introduction

During long-term service, structures may suffer damage due to material aging, corrosion, prolonged loading, and natural disasters such as fire and earthquakes [1]. For major engineering structures, such as large-span bridges, high-rise buildings, large-span spatial lattice structures, large-scale water conservancy structures, and large-scale offshore platforms, failure can cause significant economic damage and human casualties. Therefore, SHM is of great significance in civil engineering and related fields [2].
Damage identification and localization are crucial processes in SHM since damage alters the stiffness, mass, or damping of structures. Consequently, stiffness reduction caused by cracking or loosening of connections will lead to changes in structural vibration modes, resulting in alterations in its dynamic response [3,4]. However, obtaining complete and accurate high-order modal information through theoretical analysis for actual structures, particularly for large-span spatial structures with numerous degrees of freedom, is challenging due to the complex vibration data involved. In addition, in practical engineering, it is almost impossible to obtain input and output signals by artificially exciting the structure and accurately identifying the structural modal information.
The damage identification method based on time series does not involve modal parameter identification. This method is based on statistical theory and can represent a large amount of effective information contained in structural response data with fewer parameters. By identifying the change pattern of the system through the interrelationships within the data, the changes in the model parameters can be used as the basis for detecting the existence and location of damage [5]. The time series model is sensitive to the identification of small structural damage, with strong noise reduction ability and operability, and has great potential in theoretical research and engineering applications. In recent years, it has received widespread attention and development from scholars [6,7].
The ARMA model, as an important component of time domain methods, is often used to fit regression structures to acceleration response data and extract DSF from it. Zuo and Guo [8] proposed a nonlinear damage identification method based on the autoregressive (AR) model and Kullback–Leibler distance, which has high sensitivity to minor damage. Razavi et al. [4]. suggested a damage feature identification method based on ARMA model and residual sample power spectral density, introducing Jeffrey distance and Smith distance for damage localization, and verifying the effectiveness of the damage identification model through vibration response experiments. Zhu et al. [9] studied the correlation between AR coefficients and structural stiffness reduction, proposing a method to identify structural damage using underdetermined equations established by AR coefficients, and solving underdetermined equations using the sparse regularization method. The effectiveness of the method was verified through experiments. Chen et al. [10] proposed a method for identifying structural nonlinear damage based on ARMA model and vector space cosine similarity (VSCS), and verified the feasibility of the method through experimental studies, solving the nonlinear problems that traditional methods cannot effectively handle caused by structural damage. Zeng et al. [11] advanced a time series model based on fuzzy c-means clustering algorithm, characterizing the degree of structural damage by the change of model coefficients, and verifying the feasibility and accuracy of the method through experimental and numerical studies. The method has low computational cost and is suitable for real-time monitoring of civil engineering. Diao et al. [12] combined the AR model with cointegration in econometrics to cointegrate variables at different nodes of the offshore platform. The method employs cointegration residuals as damage indicators and utilizes X-bar control charts for structural damage identification. The effectiveness of this approach was validated through both numerical simulation and experimental studies. Hu et al. [13] discussed a nonlinear ARMA(n,m) time series structural damage assessment method based on residual algorithm, proposing that higher-order models are more sensitive to disguised outliers, and verifying the method through dynamic monitoring tests. Liu et al. [7] explored the implementation method of damage feature extraction and damage warning for structures based on the ARMA model, and verified the significant changes of DSF index before and after damage through a t-test.
After obtaining the DSF, structural damage identification is reduced to a pattern recognition problem. Machine learning, as the main method of pattern recognition, includes Gaussian processes, support vector machines, neural networks, etc. Tang et al. [14] presented a damage identification method based on the AR model and Gaussian process, introducing parameters that characterize damage location and state information, and achieving a probability output of multi-damage localization and damage severity. Sui et al. [15] obtained the feature vector by arranging the damage indicators obtained by the AR model in order, and inputting it into the support vector machine optimized by a Bayesian algorithm for structural damage identification. The effectiveness of the method in identifying structural damage location was verified through experiments. Xu et al. [16] proposed a damage identification method based on the AR model and a BP neural network for lattice shell structures, which does not require modal parameter identification and excitation information, avoids the dependence of damage identification results on the accuracy of structural finite element (FE) models, and has high damage identification accuracy.
This article introduced PCA on the basis of existing research to extract features from the residual sequence. The advantage of incorporating PCA lies in its ability to reduce the dimensionality of the data while retaining the most informative features [17]. This not only simplifies the subsequent analysis but also helps to identify the most significant damage indicators. In addition, the MD was used in constructing DSF because it takes into account the differences and correlations in the variability of each observed variable. In addition, it can effectively eliminate the influence of scale on different measurement values.
Furthermore, existing studies were limited to theoretical research and simple experimental verification, lacking engineering applications and practices for large and complex structures. Based on the SHM project for a timber lattice shell dome in Taiyuan Botanical Garden, this paper proposes a damage identification method based on time series analysis. The effectiveness of the damage identification model was verified using numerical simulation, and the degree of damage to the lattice shell structure was identified using measured data from SHM. This paper combined theoretical research with engineering applications to verify the effectiveness and feasibility of applying the time-series-analysis-based damage identification method for practical engineering.

2. Research Background

This study took the timber lattice shell of Taiyuan Botanical Garden as the research background, which is located in Taiyuan, Shanxi, China. It is one of the largest non-triangular timber lattice structures worldwide. It was designed by Delugan Meissl Associated Architects (DMMA) from Austria and is now a landmark building in Taiyuan. The dome, from above, is shaped like a shell and glazed with double-curved panes of glass. The span of the dome is 89.5 m, with a span-to-height ratio of 3.0 and a projected area of approximately 6000 m2. The dome consists of double-curved laminated timber beams, which are arranged in two or three intersecting layers. The beams in the intersection joint area are connected by steel pins and self-tapping screws. The splicing connection of the beam adopts the half-lap joint form and is connected by self-tapping screws. To improve the overall rigidity of the structure, a bidirectional prestressed steel cable net was installed inside it, arranged diagonally to the timber beams. The steel support of the dome is connected to the timber beam using anti-shear steel plates and self-tapping screws and connected to the concrete foundation using chemical anchor bolts. The appearance and structural details of the dome are shown in Figure 1. For more details of the dome, please refer to reference [18].

3. Methodology

This section provides an overview of the fundamental principles of utilizing the ARMA model for time series analysis. It also elucidates the techniques for determining the order and estimating the parameters of the ARMA model using AIC and LAR. Furthermore, this section offers a theoretical exposition of the principal component analysis method introduced and defines a DSF construction method based on MD.

3.1. Time Series Analysis Modeling

Time series analysis is a statistical method for dynamic data processing, which is based on the theory of stochastic processes and mathematical statistics. It uses parameter models to process ordered random sampling data and studies the statistical laws followed by data sequences, in order to perform system identification [19,20]. Any structure can be regarded as a mechanical system composed of stiffness, mass, and damping matrices, which can be described by a motion differential equation and can be transformed into a difference equation in the discrete time domain. This difference equation has the same expression form as the time series model. Therefore, the equivalent relationship between the time series model and the structural motion differential equation can be used to determine the state of the structure and perform damage identification of the system.

3.1.1. Basic Theory of ARMA Model

The ARMA model is a statistical model commonly used in time series analysis. It takes past observations of the time series as independent variables and current observations as the dependent variables to describe the randomness and autocorrelation of the time series data. The ARMA model consists of two parts: the AR model and the moving average (MA) model, and has the characteristics of both models. The AR or MA model is a special case of the ARMA model, so the ARMA model is more general for system response and is the most commonly used model in time series analysis.
For a multi-degree-of-freedom system, its vibration differential equation is:
M x ¨ t + C x ˙ t + K x t = F ( t ) ,
where M, C, and K represent the mass matrix, damping matrix, and stiffness matrix of the structure, respectively. x ¨ ( t ) , x ˙ ( t ) , and x ( t ) represent the acceleration vector, velocity vector, and displacement vector of the structure, respectively. F ( t ) is the excitation force applied to the system.
Equation (1) is equivalent to a 2n-order non-homogeneous differential equation system. Under the action of a single excitation f ( t ) , the vibration differential equation is expressed as follows:
φ 2 n x 2 n + φ 2 n 1 x 2 n 1 + + φ 1 x ˙ + φ 0 x = θ 2 n 2 f t 2 n 2 + θ 2 n 1 f t 2 n 1 + + θ 1 f t ˙ + θ 0 f t
Discretizing Equation (2) and using a white noise sequence a t ~ N D ( 0 , σ a 2 ) as the system input yields a single degree of freedom ARMA (p,q) model [21]:
x t i = 1 p φ i x t i = θ 0 a t j = 1 q θ j a t j
In the equation, φ i represents the i-th order AR coefficient, θ j represents the j-th order MA coefficient, and p and q are the AR and MA orders, respectively. By introducing a backward shift operator B, defined as B k x t = x t k [5], Equation (3) can be transformed into:
x t = 1 j = 1 q θ j B j 1 i = 1 p φ i B i a t = θ ( B ) φ ( B ) a t
According to Equation (4), the ARMA model describes a system with a transfer function of θ ( B ) / φ ( B ) , where φ ( B ) represents the inherent characteristics of the system and θ ( B ) represents the relationship between the system and the external environment. As the input signals of large civil engineering structures are often difficult to test or measure, the performance and behavior of the structure can only be inferred through the analysis of output signals. The ARMA model can be established by analyzing the output signal of the structure, without considering the specific information of the input signals, and only using the white noise sequence { a t } as input to establish an analysis model. This makes the ARMA model increasingly widely used in the field of civil engineering. In the analysis of time series models, the AR, MA, and ARMA models have their own characteristics, which can be used to preliminarily determine the appropriate time series model. The characteristics of the models are shown in Table 1.

3.1.2. Determination of ARMA Order

When applying the ARMA model for time series analysis, it is necessary to determine the order of the ARMA model, and the quality of the order setting has a significant impact on the parameter identification results. If the order is too high, the model will be too complex, resulting in overfitting and unnecessary calculations; if the order is too low, the model will be too simple to capture important features of the time series, leading to poor model fitting. Currently, one of the mostly used methods for determining the order of the ARMA model is AIC [22]. The function of AIC is defined as follows:
A I C p + q = ln σ a 2 + 2 ( p + q ) / N
In Equation (5), N represents the length of the time series x t , σ a 2 represents the variance of the ARMA model residuals, and p and q represent the AR and MA order, respectively. As the sum of model orders ( p + q ) increases, ln ( σ a 2 ) will decrease and 2 ( p + q ) / N will increase. Therefore, under the given parameter estimation method, the order with the minimum A I C p + q value should be chosen.

3.1.3. Estimation of ARMA Parameter

This article used the long autoregressive (LAR) method to estimate the parameters of the ARMA model. Compared with other methods for parameter estimation, the LAR method can transform nonlinear regression problems into linear regression problems, which can simplify the calculation process and improve efficiency. According to the theory of LAR method, the ARMA model and the AR model are equivalent mathematical models. Therefore, the parameters of the AR model can be estimated first, and then the parameters of the ARMA model can be estimated based on the relationship of the transfer functions. The equivalent relationship between the transfer functions of AR and ARMA models is as follows:
1 1 i = 1 p I i B i = 1 j = 1 m θ j B j 1 k = 1 n φ i B k
In Equation (6), the left side represents the transfer function of an AR(p) model, and the inverse function I i is equal to the parameter φ i of the AR model. The right side represents the transfer function of an ARMA(n,m) model. By moving the same power coefficients of the shift operator B in Equation (6), it can be deduced that
φ 1 = θ 1 + I 1 φ 2 = θ 2 θ 1 I 1 + I 2 φ n = θ m I n m θ 2 I n 2 θ 1 I n 1 + I n 0 = θ m I k m θ 2 I k 2 θ 1 I k 1 + I k
According to Equation (7), the solution for φ i ( i = 1,2 , , n ) can be obtained as follows:
φ 1 φ 2 φ 3 φ n = θ 1 θ 2 θ 3 θ n + 1 0 0 0 θ 1 1 0 0 θ 2 θ 1 1 0 θ n 1 θ 1 θ 1 1 I 1 I 2 I 3 I n ( when   j > 0 ,   θ j = 0 )  
For the last term in Equation (7), let k = n + 1 , n + 2 , , n + m , where n + m = p . Then, I i ( i = n + 1 , n + 2 , , n + m ) can be expressed as
I n + 1 I n + 2 I n + 3 I n + m = I n I n 1 I n 2 I n + 1 m I n + 1 I n I n 1 I n + 2 m I n + 2 I n + 1 I n I n + 3 m I n + m 1 I n + m 2 I n + m 3 I n θ 1 θ 2 θ 3 θ n
Both Equations (8) and (9) are linear equation sets about θ j , so θ j can be solved first according to Equation (9), and then φ i can be solved according to Equation (8).
Based on the theoretical research of Ljung [23], this paper adopted the order scheme of p = 2 m , q = 2 m 1 to establish the ARMA model of a multi-degree-of-freedom system, which is also a widely used model scheme in engineering [24].

3.1.4. Principal Component Analysis

There is a large amount of environmental noise in the on-site measurement data of a structure, which greatly reduces the efficiency and reliability of damage monitoring. Therefore, it is necessary to conduct feature analysis on the data to remove the noise influence and amplify the impact of structural damage on data changes. In this paper, PCA was adopted for feature extraction, which can filter out noise and redundancy in the original high-dimensional feature space data and transform it into interpretable data in a low-dimensional feature space. It can eliminate confusing data while preserving the main information [25].
For a dataset X R n × p with p variables and n samples,
X = x 11 x 12 x 1 p x 21 x 22 x 2 p x n 1 x n 2 x n p = X 1 X 2 X p
the i-th principal component of X can be represented as
Y i = u 1 i X 1 + u 2 i X 2 + + u p i X p ( i = 1 ,   2 ,   ,   p )
where u p i represents the characteristic vector. After standardizing the original data matrix X, the resulting matrix X* has a covariance matrix S that is equal to the correlation coefficient matrix:
S = 1 n 1 X * X * T
The eigenvalues of the covariance matrix of X* are denoted by λ 1 ,   λ 2 ,   , λ p , and λ 1 λ 2 λ p > 0 . The number of principal components n is determined by the variance accumulative contribution rate (ACR) i = 1 n λ i / i = 1 p λ i . If the ACR of the first n principal components is not less than 85%, it indicates that the first n principal components have incorporated most of the original data information [26].

3.2. Damage Sensitive Feature

DSF refers to the characteristic or indicators used to identify and locate potential structural damage, which can reflect the internal state changes of the structure or system. It requires the establishment of an ARMA model to analyze the signal, and the parameters of the model include the inherent characteristics of the system. By extracting the carrier containing structural features, it can be determined whether the structure is damaged. Because the AR coefficients reflect the dynamic characteristics of the system, they can be used to construct the DSF.

3.2.1. Mahalanobis Distance

The principle of the structural damage detection method based on statistical pattern recognition is to compare two sets of model parameters, one set is the model feature parameters of the structure in the healthy state, and the other set is the feature parameters in the unknown (to be identified) state. In the damage identification method based on the ARMA model, the MD is usually used to determine the distance between the sample data and the reference data after feature extraction, and then evaluate the health status of the structure. MD calculates the distance between two multivariate sets by considering the correlation between them, which is an effective method for computing the similarity between two unknown-state sample sets. The expression of the MD between the sample set z and the population G is:
d M ( z , G ) = ( z v G ) T 1 ( z v G ) 1 2
where v G represents the mean vector of G, and Σ represents the covariance matrix of G.

3.2.2. Construction of DSF

ARMA models were established for the unknown-state sample set, reference sample set, and training sample set. The DSF was constructed by measuring the differences between the data parameters of the ARMA model under the unknown state and the healthy state. The formula is as follows:
D S F ( u , r ) = d M 2 ( u , T ) d M 2 ( r , T )
where u represents the unknown-state sample set, r represents the reference sample set of the healthy state, and T represents the training sample set of the healthy state. A DSF value close to 1 indicates that the structure is in a healthy state, while a DSF value greater than 1 indicates damage. Moreover, the severity of the damage is positively correlated with the magnitude of the DSF.

4. Results and Discussions

This section provides an explanation and discussion of the damage identification results using the FE models and SHM measured data. Firstly, a concise introduction is provided for the FE model, which includes the setup of five simulated DCs. The establishment of a time series model and the detailed results of damage identification are then elaborated. Finally, the obtained results are thoroughly explained and discussed. Furthermore, this section is based on two years of SHM data collected from the timber lattice shell dome of Taiyuan Botanical Garden. The proposed damage identification method from this article is applied to identify the damage in the dome, and the identification results are subsequently analyzed and discussed.

4.1. Damage Identification Based on Numerical Models

Structural damage in practical engineering exhibits characteristics such as randomness and finiteness, making it difficult to determine the effectiveness and availability of time series models through measured data. However, FE methods can be used to simulate the health and damage status of the structure, in order to establish and validate the ARMA model.
To verify the feasibility and applicability of the ARMA model for large-span timber lattice structures, this paper used Midas Gen, a general structural analysis software, to perform FE modeling and analysis for the timber lattice dome of Taiyuan Botanical Garden. The stiffness of individual or multiple components was reduced to simulate the structural damage conditions (DC). Gaussian white noise was used to excite the structure, and the response results of multiple measuring points (MP) were extracted to evaluate the damage identification effect of the ARMA model.

4.1.1. Finite Element Model

This paper used 3D line elements to analyze the spatial structure of the dome. The timber beams were modeled with beam elements, while the cable net was modeled with cable (tensile only) elements. In this study, the maximum main stress and maximum vertical deformation were employed as the convergence indicators. The analysis findings demonstrated that as the number of elements increased from 7732 to 334,817, both the maximum stress and maximum deformation exhibited a gradual convergence. The mesh convergence plot is shown in Figure 2. Notably, the maximum stress was more sensitive to the element number, thus serving as the primary controlling factor. Taking into account both computational accuracy and efficiency, the 134,372-element model was ultimately selected, which achieved a maximum stress convergence within a 5% range and satisfied the required computational accuracy.
The FE model took into account the fact that the three-layer timber beams are not coplanar, as shown in Figure 3. The timber beams are made of European spruce glued laminated timber GL28h, with a main beam section of 200 mm × 400 mm (width × height) and a secondary beam section of 200 mm × 300 mm (width × height). The components of the lattice structure are mainly subjected to axial forces, and glued laminated timber has relatively ideal material properties. According to material tests [27] and related studies [28,29,30], the FE material model adopted an orthotropic bilinear elastic-plastic model. The material test data and FE material model were shown in Figure 4, and the material properties are detailed in Table 2. The cable has a diameter of 26 mm, and the material used is austenitic steel 06Cr17Ni12Mo2 (316 stainless steel), with an elastic modulus of 160,000 MPa and a design pre-tension value of 40 kN.
Based on the measured data of SHM, simple modifications were attempted on the FE model. Initially, the acceleration signal during the early stage of structural service was processed and analyzed. The time-domain data were then transformed into frequency-domain data using Fourier transform. To eliminate high-frequency noise, a low-pass filter of 50 Hz was applied. The power spectral density (PSD) of the acceleration data was calculated, and the dominant frequency with significant amplitude in the signal was extracted through peak detection. This allowed for the determination of the natural frequency of the structure.
Subsequently, modifications were made to the FE model parameters. The elastic modulus and density of the material were adjusted, along with the magnitude of the additional constant load (converted into mass) on the roof, and the connection stiffness of the joints. These modifications aimed to align the natural vibration characteristics of the FE model with the analysis results obtained from the measured data.

4.1.2. Damage Conditions

When establishing a time series model, the first step is to determine the sampling time interval Δ t and the sample length L for the continuous signal. When sampling a continuous signal, a reasonable sampling time interval is Δ t < 1 / ( 2 f m a x ) , and a reasonable sample length is L > 1 / Δ f m i n , where f m a x is the highest frequency in the interested vibration modes and Δ f m i n is the minimum difference between adjacent frequencies [32].
Given the complex natural vibration characteristics of the lattice shell structure, a 100-order natural vibration modal analysis was conducted for the FE model. The results of the modal analysis are presented in Figure 5 and Table 3. The analysis revealed that the natural frequencies of the lower order modes are closely spaced, with a difference of only 8.92 Hz between the first and 70th modes. By the 94th mode, 90% of the cumulative participation mass of the vibration modes was reached, indicating that the first 94 modes could capture the majority of the natural vibration characteristics of the structure. Therefore, the appropriate sampling frequency and sample length were calculated based on the first 94 modes. The maximum frequency in the first 94 modes was approximately 49.85 Hz, and the minimum difference between adjacent frequencies was 0.0109 Hz. According to the sampling theories [32], the sampling time interval of this structure should be less than 0.0101 s, and the sample length should be greater than 91.75 s. Therefore, a sampling frequency of 100 Hz and a sample length of 120 s were adopted to establish a time series model. According to the basic principle of time series analysis, the signal input needs to meet the characteristics of stationary, normal, and zero-mean, and Gaussian white noise can just meet the above characteristics. Thus, Gaussian white noise generated by MATLAB was used as the environmental excitation to conduct the time history analysis on the FE model.
In practical engineering, vibration sensors were used for SHM of the dome structure. Specifically, nine vibration MPs were set up at the joint positions, taking into account the shape and characteristics of the structure. From above, the MPs were distributed at the central point of the dome and in the directions of east, west, south, and north. In the vertical direction, the MPs were evenly distributed at the top, middle, and bottom of the structure. During the time history analysis on the FE models, acceleration response data were extracted from the nine MPs. The layout of the MPs was shown in Figure 6. In the FE analysis process, stiffness reduction was applied to local timber beam and steel cable elements to simulate damage or stiffness degradation of the components. This paper developed five sets of DCs, including damage to single and multiple components. The five simulated DCs are detailed in Table 4.

4.1.3. Establishment of ARMA Model

This study established time series models for all nine MPs under various DCs as shown in Table 4. In this paper, the process of establishing a time series model was demonstrated using the acceleration time history data of MP M, which is located at the top of the dome structure.
Before establishing the ARMA model, it is necessary to standardize the sample dataset. The normal distribution probability plot of the processed data, which was shown in Figure 7c, indicated that the processed sample is a stationary, normal, and zero-mean time series, satisfying the prerequisite for establishing a time series model. After preprocessing the data, the ACF and PACF of the data sample were calculated, as shown in Figure 7a,b. The ACF and PACF gradually decreased to within the 95% confidence interval, and both exhibited tailing characteristics. According to Table 1, an ARMA model could be established for the data sample.
Subsequently, the AIC and LAR method were used to determine the order and estimate the parameters of the model. Figure 7d shows the AIC results of the ARMA(2m,2m−1) model for the first 30 orders. Based on the AIC results, an ARMA(20,19) model was established for the MP M, and its estimated parameters are shown in Table 5.

4.1.4. Damage Identification of FE Models

ARMA(p,q) models were established for the responses obtained from the FE models under the five DCs shown in Table 4. To alleviate the computational burden arising from the high AR order, dimensionality reduction could be achieved by conducting PCA on its parameters. Based on the principle of PCA, the AR parameters of each MP were analyzed for five distinct DCs.
Table 6 presents the eigenvalue and ACR obtained through PCA for the data of each MP. The eigenvalue quantifies the extent to which each principal component explains the variation in the data, while the ACR represents the proportion of total variance explained by the first n principal components. An ACR exceeding 85% indicates that the primary components already encompass most of the feature information [26]. Based on the PCA results, the eigenvalues of the first and second order AR coefficients for each MP were nearly 10 and 4, respectively. Furthermore, the ACRs for the first three orders at each MP all exceeded 90%, surpassing the 85% threshold. In some cases, the ACR even reached approximately 99%. Consequently, it can be inferred that employing the first three order AR coefficients to construct DSF can yield sufficiently accurate outcomes in theory.
The acceleration time series data from each MP under the healthy state (DC 1) was subjected to windowing processing, resulting in two equally divided sections. The first 60 s of the resulting time series data, comprising 6000 sample data points, was utilized as training sample data, while the last 60 s, also comprising 6000 sample data points, was used as reference sample data. The remaining data obtained from other DCs were unknown-state sample data. The first three order AR coefficients of the ARMA model were extracted for each sample data. A principal component matrix was established and the MD was calculated between the principal component matrices of a single MP under different DCs. Corresponding DSF for damage identification was constructed according to Equation (14). The DSF for each MP is presented in Figure 8.
Based on the results of damage identification, the following observations could be made:
(1)
When the damage of the timber beam at MP M was 40% (DC 2), the DSF at M was 2.1799, which was significantly higher than the DSF at other MPs, not exceeding 1.15. As the damage increased to 80% (DC 3), the DSF at M increased to 3.5720, while the average DSF at other MPs was 1.1450. This represented a 63.9% increase in DSF compared to the 40% damage level. When both the timber beam and steel cable at M were damaged simultaneously (DC 4), the DSF further increased to 4.0281, which was approximately 12.77% higher than the previous DC. The average DSF value at other MPs was 1.1773. These results demonstrate that the DSF constructed using PCA and MD is highly sensitive to component damage, capable of identifying and locating damage to timber beams and steel cables, and positively correlates with the degree of damage. Therefore, it could reflect the degree of damage to the damaged components to a certain extent.
(2)
When the timber beam and steel cable at MPs M and N1 were simultaneously damaged by 80% (DC 5), the DSF of M and N1 were 4.1645 and 3.9983, respectively, which were significantly higher than the DSF values of other MPs, ranging from 1.07 to 1.38. This demonstrates that the proposed damage identification model can effectively identify conditions where multiple components are damaged simultaneously.
(3)
As the damage at MP M intensified from DC 2 to 4, the DSF of M significantly increased, while the DSF of adjacent MPs N1, E1, S1, and W1 also slightly increased, with an average increase of approximately 0.19. However, the DSF of MPs N2, E2, S2, and W2 did not increase significantly, with an average increase of no more than 0.1. Under DC 5, further damage occurred at MP N1. During this process, the DSF of N2 and M, which were closest to N1, increased the most, with values of 0.2070 and 0.1364, respectively. However, the DSF of S2, W2, and E2, which were farthest from N1, remained almost unchanged, with values of 0.049, 0.074, and 0.054, respectively. This suggests that structural damage at a specific location can cause an increase in the DSF of surrounding MPs, and the degree of increase is inversely proportional to the distance. This phenomenon may be attributed to the changes in the local natural vibration characteristics caused by component damage, which were not significantly far from the damage location.
(4)
Assuming that no MPs are set at M and N1, meaning that the DSF of M and N1 were not taken into account, under DC 4, the DSF values of E1, W1, and S1 experienced a significant increase, ranging from 1.25 to 1.35, while the DSF of the remaining MPs showed no noticeable changes. This suggests that the damage occurred in areas near E1, W1, and S1. It can be inferred that the damage may have occurred around M, which aligns with the severe damage observed at M under DC 4. From DC 4 to 5, the DSF of N2 increased significantly, while the DSF of the other MPs increased slightly and to a similar extent. This implies that the damage may have occurred in a location close to N2. As determined, N1 sustained damage during the transition from DC 4 to 5, which is consistent with this characteristic. The aforementioned statement indicates that even if the damage does not occur at MPs, the approximate location of the damage can be determined based on the DSF of the existing MPs, thereby achieving the objective of damage monitoring. This is advantageous for optimizing the number of MPs.

4.2. Damage Identification Based on SHM Data

This paper presented a damage identification study of the dome structure at Taiyuan Botanical Garden, utilizing measured data from vibration sensors. The structure employed glued laminated timber beam as its primary load-bearing component, and exhibited a large span, complex natural vibration characteristics, and susceptibility to wind vibration, earthquakes, and other impacts. To ensure the structure’s safety, long-term SHM has been conducted since its service. Based on the structure’s characteristics and FE analysis results, nine vibration MPs shown in Figure 6 were selected to monitor the structure’s vibration behavior. These MPs were equipped with the 2D001 magnetoelectric vibration sensors produce by Donghua Testing Technology Co. Ltd. (Taizhou, China), with a maximum sampling frequency of 100 Hz. The data acquisition device is shown in Figure 9, and the technical parameters of the 2D001 sensor are shown in Table 7.
Under normal working conditions, the only dynamic load on this structure was wind load. However, due to the fact that the dome is covered with glass, the structure was exposed to sunlight radiation during the day, resulting in temperature gradients on the surface that may have an uncertain impact on the sampling data and final analysis results. Therefore, this study selected measured data at night as the sample data. Specifically, acceleration data from the initial stage of structural service (October 2020) were chosen as the healthy-state sample data, while data after two years of service (October 2022) were selected as the unknown-state sample data, with a sample length of 120 s. Hence, ARMA modeling was performed and the health status of the structure after two years of service was evaluated. Figure 10 displays partial data for the initial stage and after two years of service.
The results of the calculation indicated that the ACF and PACF of the model are tailing, which satisfies the prerequisite for ARMA modeling. The AIC was employed to determine the order of the ARMA(2m,2m−1) model for the measured data. The findings reveal that the ARMA(28,27) model had the lowest AIC value of −1.4235 among the first 30 orders, indicating that this order scheme can achieve a high degree of fitting. The vertical acceleration data of nine vibration sensors at each MP were preprocessed, and ARMA(28,27) models were established. PCA was performed on the AR parameters using a set of measured data from each MP under both healthy and unknown states. The analysis results showed that the first 5 order AR parameters of any ARMA model contained 90% of the model’s damage characteristic values. Therefore, the first 5 order AR coefficients were utilized to construct subsequent DSF.
The acceleration time series of each MP in a healthy state was extracted and partitioned into two equal segments by windowing. The first 60 seconds’ 6000 sample data points were designated as training sample data, while the last 60 seconds’ 6000 sample data points were designated as reference sample data. The remaining data were used as unknown-state sample data. The MD was computed to generate DSF for damage identification.
According to the analysis results, the MDs of each sample were not significantly different, and the DSF of all MPs were close to 1.0, indicating that the structure was almost not damaged after two years of service. The results of MD and DSF of each MP were shown in Figure 11. In addition, it could be found that the DSF of MP S2 was 1.097, while the maximum value of DSF of other MPs was 1.062, and the average value was 1.046, indicating that the damage degree of the component at S2 was slightly higher than that of other MPs. Figure 12 shows the static FE analysis results of the dome. Under the combined action of dead load and roof live load, the stress level of the timber components around S2 was higher than that in other areas. Moreover, real-time dynamic loads such as wind loads and temperature effects might have adverse effects on local areas. These adverse effects might be the potential factors leading to the increase of DSF at S2. Therefore, it is necessary to pay attention to the SHM data of this dome structure, and focus on the acceleration data of MP S2, so as to timely warn when the structure suffers serious damage.

5. Conclusions

Structural damage identification is a crucial analysis method in SHM, with significant scientific and engineering applications. To address the challenges of incomplete modal information, high modal density, and large degrees of freedom in large-span spatial lattice structures, this paper proposed a structural damage identification method based on the ARMA model and PCA method for structures of this kind. Based on the above research, the following conclusions can be drawn:
(1)
This article was based on a time series model and introduced PCA for extracting principal components. In this study, only the first three AR coefficients were required to achieve damage identification and it demonstrated high accuracy.
(2)
The MD is advantageous compared to other distance methods in constructing DSF because it considers the differences and correlations in the variability of variables for each observation. Additionally, it can effectively calculate the influence of scale on different measurement values.
(3)
Verified by the FE model, the damage identification method proposed in this article was found to be sensitive to structural damage. It could, to a certain extent, accurately locate the site of structural damage and reflect the extent of the damage. Therefore, it can be effectively utilized for practical structural damage identification.
(4)
Based on SHM data, the paper identified the damage of the structure after two years of service. The results indicated that the structure is in a relatively healthy state, but the DSF of MP S2 is slightly higher than those of other MPs. In the future, it is necessary to pay attention and give timely warnings if necessary.
However, there are still some shortcomings in this article. The next step should be to conduct more in-depth research on the rationality of the number and distribution of measurement points, as well as the refinement of the FE model correction.

Author Contributions

Conceptualization, G.W. and H.L.; methodology, G.W., C.X. and H.L.; software, G.W. and C.X.; validation, G.W. and S.Z.; formal analysis, G.W., C.X. and S.Z.; investigation, G.W., C.X., S.Z. and Z.Z.; data curation, Z.Z. and L.Z.; writing—original draft preparation, G.W.; writing—review and editing, H.L.; visualization, J.W. and B.Q.; supervision, H.L.; project administration, L.Z. and B.Q.; funding acquisition, C.X. and L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant number 52278198; and by the Structural Health Monitoring Project for Taiyuan Botanical Garden, grant number 0632-1920FW3L2061-01.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available from the first author, ([email protected]), upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhao, Y.; Gong, M.; Yang, Y. A review of structural damage identification methods. World Earthq. Eng. 2020, 36, 73–84. [Google Scholar]
  2. Yan, G.; Duan, Z.; Ou, J. Review on structural damage detection based on vibration data. J. Earthq. Eng. Eng. Vib. 2007, 27, 95–103. [Google Scholar] [CrossRef]
  3. Lakshmi, K.; Rao, A.R.M. Damage identification technique based on time series models for LANL and ASCE benchmark structures. Insight-Non-Destr. Test. Cond. Monit. 2015, 57, 580–587. [Google Scholar] [CrossRef]
  4. Razavi, B.S.; Mahmoudkelayeh, M.R.; Razavi, S.S. Damage Identification under Ambient Vibration and Unpredictable Signal Nature. J Civ. Struct Health Monit 2021, 11, 1253–1273. [Google Scholar] [CrossRef]
  5. Box, G.E.P.; Jenkins, G.M. Time Series Analysis, Forecasting and Control; John Wiley & Sons: Hoboken, NJ, USA, 1971; Volume 134. [Google Scholar]
  6. Bao, C.; Hao, H.; Li, Z.-X. Integrated ARMA Model Method for Damage Detection of Subsea Pipeline System. Eng. Struct. 2013, 48, 176–192. [Google Scholar] [CrossRef]
  7. Liu, Y.; Li, A.; Ding, Y.; Fei, Q. Time Series Analysis with Structural Damage Feature Extraction and Alarming Method. Chin. J. Appl. Mech. 2008, 25, 253–257+357. [Google Scholar]
  8. Zuo, H.; Guo, H. Structural Nonlinear Damage Identification Method Based on the Kullback-Leibler Distance of Time Domain Model Residuals. Remote Sens. 2023, 15, 1135. [Google Scholar] [CrossRef]
  9. Zhu, H.; Yu, H.; Gao, F.; Weng, S.; Sun, Y.; Hu, Q. Damage Identification Using Time Series Analysis and Sparse Regularization. Struct. Control Health Monit. 2020, 27, e2554. [Google Scholar] [CrossRef]
  10. Chen, L.; Yu, L.; Fu, J.; Ng, C.-T. Nonlinear Damage Detection Using Linear ARMA Models with Classification Algorithms. Smart Struct. Syst. 2020, 26, 23–33. [Google Scholar] [CrossRef]
  11. Zeng, Y.; Yan, Y.; Weng, S.; Sun, Y.; Tian, W.; Yu, H. Fuzzy Clustering of Time-Series Model to Damage Identification of Structures. Adv. Struct. Eng. 2019, 22, 868–881. [Google Scholar] [CrossRef]
  12. Diao, Y.; Cao, Y.; Sun, Y. Structural Damage Identification based on AR Model Coefficients and Cointegration for Offshore Platform Under Environmental Variations. Eng. Mech. 2017, 34, 179–188. [Google Scholar]
  13. Hu, M.-H.; Tu, S.-T.; Xuan, F.-Z. Statistical Moments of ARMA(n,m) Model Residuals for Damage Detection. Procedia Eng. 2015, 130, 1622–1641. [Google Scholar] [CrossRef]
  14. Tang, Q.; Xin, J.; Zhou, J.; Fu, L.; Zhou, B. Structural damage identification method based on AR-GP model. J. Vib. Shock 2021, 40, 102–109. [Google Scholar] [CrossRef]
  15. Sui, Z.; Wang, Q.; Jia, D.; Yang, Y. Structural Damage Identification based on AR Model and Bayesian Optimization SVM. Low Temp. Archit. Technol. 2020, 42, 74–76. [Google Scholar] [CrossRef]
  16. Xu, J.; Guo, W.; Wang, X.; Cui, S. Study on the damage detection method for latticed shell structures based on AR model and BP neural network. Spat. Struct. 2015, 21, 66–71. [Google Scholar] [CrossRef]
  17. Wang, C.; Guan, W.; Gou, J.; Hou, F.; Bai, J.; Yan, G. Principal Component Analysis Based Three-Dimensional Operational Modal Analysis. Int. J. Appl. Electromagn. Mech. 2014, 45, 137–144. [Google Scholar] [CrossRef]
  18. Jia, S. Research on key technologies of large-span glued laminated timber reticulated shell structure design in Taiyuan Botanical Garden. Build. Struct. 2022, 52, 1–5+62. [Google Scholar] [CrossRef]
  19. Liu, Z. Spatial Steel Structural Damage Alarming Based on Statistical Pattern Recognition. Master’s Thesis, Harbin Institute of Technology, Harbin, China, 2010. [Google Scholar]
  20. Zhou, X. Research on the Damage Identification of Girder Bridge Structures Based on Auto-Regressive Moving Average Model. Master’s Thesis, Southwest Jiaotong University, Chengdu, China, 2008. [Google Scholar]
  21. Du, Y.; Li, W.; Li, H.; Liu, D. Structural damage identification based on time series analysis. J. Vib. Shock 2012, 31, 108–111. [Google Scholar] [CrossRef]
  22. Burnham, K.; Anderson, D. Model Selection and Multimodel Inference: A Practical Information-Theoretic Approach, 2nd ed.; Springer: Berlin/Heidelberg, Germany, 2002; Volume 67, ISBN 978-0-387-95364-9. [Google Scholar]
  23. Ljung, L. System Identification: Theory for the User. In System Identification: Theory for the User; Prentice-Hall: Hoboken, NJ, USA, 1999. [Google Scholar]
  24. Yang, S. Time Series Analysis in Engineering Application, 2nd ed.; Huazhong University of Science & Technology Press: Wuhan, China, 1991. [Google Scholar]
  25. Zhou, Y.; Liu, Z.; Zhang, G.; Zhou, L.; Liu, Y.; Tang, L.; Jiang, Z.; Yang, B. Structural Damage Identification Method Based on Moving Principal Component Analysis and Ensemble Learning. J. Univ. Jinan Sci. Technol. 2023, 37, 116–126. [Google Scholar] [CrossRef]
  26. Kumar, K.; Biswas, P.K.; Dhang, N. Time Series-Based SHM Using PCA with Application to ASCE Benchmark Structure. J. Civ. Struct. Health Monit. 2020, 10, 899–911. [Google Scholar] [CrossRef]
  27. Jia, S.; Li, R.; Li, Y. Research on material properties test and application of imported glued laminated timber in Taiyuan Botanical Garden. Build. Struct. 2022, 52, 17–21. [Google Scholar] [CrossRef]
  28. Zhang, L.; Xie, Q.; Wu, Y.; Zhang, B. Research on the mechanical properties and constitutive model of wood under parallel-to-grain cyclic loading. China Civ. Eng. J. 2023, 44, 286–304. [Google Scholar]
  29. Wang, M.; Song, X.; Gu, X. Nonlinear analysis of wood based on three-dimensional combined elastic-plastic and damage model. China Civ. Eng. J. 2018, 51, 22–28+49. [Google Scholar] [CrossRef]
  30. Wang, M.; Gu, X.; Song, X.; Tang, J. State-of-the-art for description methods of nonlinear mechanical behavior of wood. J. Build. Struct. 2021, 42, 76–86. [Google Scholar] [CrossRef]
  31. Xu, B.-H.; Bouchaïr, A.; Racher, P. Appropriate Wood Constitutive Law for Simulation of Nonlinear Behavior of Timber Joints. J. Mater. Civ. Eng. 2014, 26, 04014004. [Google Scholar] [CrossRef]
  32. Shannon, C.E. Communication in the Presence of Noise. Proc. IEEE 1998, 86, 447–457. [Google Scholar] [CrossRef]
Figure 1. Timber lattice shell dome of Taiyuan Botanical Garden. (a) Appearance; (b) joint and its spatial relationship; (c) bidirectional cross steel cable; (d) root support.
Figure 1. Timber lattice shell dome of Taiyuan Botanical Garden. (a) Appearance; (b) joint and its spatial relationship; (c) bidirectional cross steel cable; (d) root support.
Sensors 23 06710 g001aSensors 23 06710 g001b
Figure 2. Mesh convergence analysis.
Figure 2. Mesh convergence analysis.
Sensors 23 06710 g002
Figure 3. FE model. (a) Overall structure; (b) timber beam’s stacking joint.
Figure 3. FE model. (a) Overall structure; (b) timber beam’s stacking joint.
Sensors 23 06710 g003
Figure 4. Material properties. (a) Material test data (GL28h), reprinted with permission from Ref. [27]. 2022, Shuizhong Jia; (b) bilinear material model of glued laminated timber [31].
Figure 4. Material properties. (a) Material test data (GL28h), reprinted with permission from Ref. [27]. 2022, Shuizhong Jia; (b) bilinear material model of glued laminated timber [31].
Sensors 23 06710 g004
Figure 5. 100-order Z-axis modal analysis.
Figure 5. 100-order Z-axis modal analysis.
Sensors 23 06710 g005
Figure 6. Distribution of MPs and damaged components.
Figure 6. Distribution of MPs and damaged components.
Sensors 23 06710 g006
Figure 7. ARMA order selection. (a) ACF; (b) PACF; (c) normal distribution probability; (d) AIC of each order combination.
Figure 7. ARMA order selection. (a) ACF; (b) PACF; (c) normal distribution probability; (d) AIC of each order combination.
Sensors 23 06710 g007
Figure 8. Damage identification results of FE models. (a) Results under DC 2; (b) results under DC 3; (c) results under DC 4; (d) results under DC 5; (e) results of M and N1; (f) results of undamaged MPs.
Figure 8. Damage identification results of FE models. (a) Results under DC 2; (b) results under DC 3; (c) results under DC 4; (d) results under DC 5; (e) results of M and N1; (f) results of undamaged MPs.
Sensors 23 06710 g008
Figure 9. Data acquisition device. (a) Tri-dictional vibration monitoring module; (b) 2D001 magnetoelectric vibration sensor.
Figure 9. Data acquisition device. (a) Tri-dictional vibration monitoring module; (b) 2D001 magnetoelectric vibration sensor.
Sensors 23 06710 g009
Figure 10. Original SHM data (M and N1).
Figure 10. Original SHM data (M and N1).
Sensors 23 06710 g010
Figure 11. Damage identification results. (a) MD results; (b) DSF results.
Figure 11. Damage identification results. (a) MD results; (b) DSF results.
Sensors 23 06710 g011
Figure 12. High stress area exhibited in static FE analysis result.
Figure 12. High stress area exhibited in static FE analysis result.
Sensors 23 06710 g012
Table 1. Basic characteristics of time series models.
Table 1. Basic characteristics of time series models.
Model TypeAutocorrelation
Function (ACF)
Partial Autocorrelation
Function (PACF)
ARTailingTruncation
MATruncationTailing
ARMATailingTailing
Table 2. Material property of GL28h.
Table 2. Material property of GL28h.
Ec,0 (MPa)Et,0 (MPa)Ec,90 (MPa)fc,0 (MPa)ft,0 (MPa)fc,90 (MPa)
13,901743818529.677.55.0
Table 3. Natural frequencies of the FE model.
Table 3. Natural frequencies of the FE model.
Mode
Order
Frequency
(Hz)
Mode
Order
Frequency
(Hz)
Mode
Order
Frequency
(Hz)
Mode
Order
Frequency
(Hz)
13.3537265.8696518.20397614.8373
23.5002275.9719528.36257715.1701
33.5525286.134538.45757815.8307
43.6825296.2517548.56357916.6591
53.7451306.2953558.72138017.2487
63.9603316.3233568.8478117.9974
74.0067326.5079579.05858219.1284
84.1308336.6175589.19138320.0353
94.2118346.6899599.37148421.7652
104.431356.7379609.53928522.4354
114.4916366.9291619.78738624.1665
124.5736376.9573629.98238726.2336
134.6526386.96826310.24078829.05
144.6659397.06496410.52358931.3453
154.9154407.23446510.66729032.7382
165.0791417.36146610.97649136.9209
175.1145427.38116711.32599240.3238
185.1646437.41716811.54949345.1509
195.2276447.55686912.06469449.8536
205.3595457.68867012.27179559.3389
215.4691467.71937112.61629672.188
225.6584477.77737213.09929788.9345
235.6804487.86687313.614998121.4361
245.7494497.97977413.986499180.6925
255.7668508.10437514.4609100349.0827
Table 4. Damage conditions.
Table 4. Damage conditions.
Condition
No.
MP MMP N1
BeamCableBeamCable
1Sensors 23 06710 i001Sensors 23 06710 i001Sensors 23 06710 i001Sensors 23 06710 i001
2Sensors 23 06710 i002Sensors 23 06710 i001Sensors 23 06710 i001Sensors 23 06710 i001
3Sensors 23 06710 i003Sensors 23 06710 i001Sensors 23 06710 i001Sensors 23 06710 i001
4Sensors 23 06710 i003Sensors 23 06710 i003Sensors 23 06710 i001Sensors 23 06710 i001
5Sensors 23 06710 i003Sensors 23 06710 i003Sensors 23 06710 i003Sensors 23 06710 i003
Note: Sensors 23 06710 i001: in a healthy state; Sensors 23 06710 i002: damaged by 40%; Sensors 23 06710 i003: damaged by 80%.
Table 5. Estimated ARMA model parameters.
Table 5. Estimated ARMA model parameters.
OrderAR Coef.MA Coef.OrderAR Coef.MA Coef.
10.19560.9294110.0487−0.3235
2−0.36410.135912−0.0731−0.1088
3−0.0754−0.396513−0.00820.3625
4−0.2056−0.401414−0.47300.8166
5−0.1359−0.0875150.10080.4343
6−0.17490.177616−0.2690−0.0291
70.24840.402317−0.1583−0.1125
8−0.14140.149518−0.31390.0210
90.0141−0.2229190.02080.1027
100.0242−0.2021200.0635-
Table 6. PCA results.
Table 6. PCA results.
Principal
Component
ME1E2
EigenvalueACREigenvalueACREigenvalueACR
110.324551.62%14.469172.35%13.732068.66%
24.630474.77%3.222588.46%3.171084.52%
33.813393.84%1.254994.73%1.921994.12%
40.855198.12%1.000299.73%1.034399.30%
50.3768100.00%0.0532100.00%0.1407100.00%
Principal
Component
W1W2S1
EigenvalueACREigenvalueACREigenvalueACR
115.056575.28%10.491252.46%12.879164.40%
24.075395.66%6.406384.49%4.029484.54%
30.750599.41%1.827393.62%2.869998.89%
40.111799.97%1.084299.05%0.215599.97%
50.0060100.00%0.1909100.00%0.0061100.00%
Principal
Component
S2N1N2
EigenvalueACREigenvalueACREigenvalueACR
110.875954.38%8.396541.98%11.138555.69%
24.356676.16%7.400978.99%7.179691.59%
32.772390.02%2.454791.26%1.104097.11%
41.942599.74%1.234097.43%0.408699.15%
50.0527100.00%0.5139100.00%0.1693100.00%
Table 7. Technical parameters of the 2D001 magnetoelectric vibration sensor.
Table 7. Technical parameters of the 2D001 magnetoelectric vibration sensor.
Mode0123
ParameterAccelerationLow VelocityMedium VelocityHigh Velocity
Sensitivity0.3 V/m·s−220 V/m·s−15 V/m·s−10.3 V/m·s−1
Capacity20 m·s−20.125 m·s−10.3 m·s−10.6 m·s−1
Bandwidth
(−3–+1 dB)
(0.25–100) Hz(1–100) Hz(0.5–100) Hz(0.17–80) Hz
Output
Impedance
50 kΩ
Working
Temperature
(−20–80) °C
Dimensions63 mm × 63 mm × 63 mm
Weight0.6 kg
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

Wang, G.; Xu, C.; Zhang, S.; Zhou, Z.; Zhang, L.; Qiu, B.; Wan, J.; Lei, H. Exploration of Damage Identification Method for a Large-Span Timber Lattice Shell Structure in Taiyuan Botanical Garden based on Structural Health Monitoring. Sensors 2023, 23, 6710. https://doi.org/10.3390/s23156710

AMA Style

Wang G, Xu C, Zhang S, Zhou Z, Zhang L, Qiu B, Wan J, Lei H. Exploration of Damage Identification Method for a Large-Span Timber Lattice Shell Structure in Taiyuan Botanical Garden based on Structural Health Monitoring. Sensors. 2023; 23(15):6710. https://doi.org/10.3390/s23156710

Chicago/Turabian Style

Wang, Guoqing, Chenjia Xu, Shujia Zhang, Zichun Zhou, Liang Zhang, Bin Qiu, Jia Wan, and Honggang Lei. 2023. "Exploration of Damage Identification Method for a Large-Span Timber Lattice Shell Structure in Taiyuan Botanical Garden based on Structural Health Monitoring" Sensors 23, no. 15: 6710. https://doi.org/10.3390/s23156710

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