Next Article in Journal
Insights on the Mortars of Ancient Roman Aqueducts: Aqua Virgo and Aqueduct Y, Rome (Italy)
Previous Article in Journal
Enhancing Thermal Efficiency through the Use of Graphite-Infused Phase Change Materials in Roof Structures to Reduce Building Cooling Demand
Previous Article in Special Issue
Advancements in Optimal Sensor Placement for Enhanced Structural Health Monitoring: Current Insights and Future Prospects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Deflection Monitoring Method for Simply Supported Girder Bridges Using Strain Response under Traffic Loads

1
College of Civil and Transportation Engineering, Hohai University, Nanjing 210098, China
2
Engineering Research Center of Intelligent Construction and Industrialization, CAAC, Tianjin 300456, China
*
Author to whom correspondence should be addressed.
Buildings 2024, 14(1), 70; https://doi.org/10.3390/buildings14010070
Submission received: 5 December 2023 / Revised: 18 December 2023 / Accepted: 23 December 2023 / Published: 26 December 2023
(This article belongs to the Special Issue Advances in Structural Monitoring for Infrastructures in Construction)

Abstract

:
Deflection measurements are usually used as a key index in civil engineering for performing structural assessments of bridge safety. However, owing to technical or cost issues, it may be difficult to implement long-term monitoring of bridge deflection, especially for short- or medium-span bridges. Therefore, this study presents a novel method for measuring the deflection of simply supported girder bridges. In the proposed method, the strain measurement was implemented under traffic loading at only one position, such as middle span, and then the strain distribution along the girder was reconstructed to calculate the girder deflection with basic structural mechanical theory. To implement the method, the theory was constructed based on the displacement reciprocal theorem at first to assess the strain distribution along the girder from the strain measurement at some position during traffic loads passing across the bridge. Second, a strain measurement method, namely long-gauge fibre Bragg grating (FBG) sensing technology, was introduced to take strain measurements for a concrete bridge. Third, various finite element (FE) bridge models were developed to validate the proposed method’s accuracy, the results from which indicated that the method accurately implemented deflection measurement with an approximately 5% calculation error. In addition, the influence of some key parameters, such as vehicle type, vehicle speed, and structural damage, was investigated. The simulation results revealed that damage to the hinge joint in the middle location could significantly influence the proposed method’s accuracy such that the error may exceed 10%. Finally, on-site experiments were conducted on a simply supported girder bridge to further validate the proposed method’s accuracy, and an approximately 8% deflection assessment error was found. Considering the additional advantages of FBG sensing technology, the proposed method can also be effective for long-term deflection measurements of short- or medium-span bridges.

1. Introduction

Simply supported girder bridges are typical types of highway bridges in China. Often referred to as hinge joints, transverse connections are used to link precast girders for vehicle load bearing [1]. However, the hinge joint may break easily under repeated traffic loads. In addition, concrete cracks, steel erosion, and concrete degradation can also influence bridge safety [2,3,4]. Therefore, effective measurement of key parameters for evaluating a bridge’s structural performance is crucial.
Deflection is often used as a key indicator for assessing the structural performance of actual bridges [5,6]. One reason is that it is easily used to implement the structural performance assessment according to the design code. Another reason is that it can be measured by simple methods, such as geodetic techniques, linear variable differential transformers (LVDT), a global positioning system (GPS), and so on [7,8].
For geodetic techniques, total station [9] and static level [10,11] are widely used in deflection measurement due to their high accuracy (up to 0.1 mm) and stability. However, they cannot be utilized to monitor dynamic deflection under traffic loads because their systems’ operation is time-consuming and laborious. Another method, linear variable differential transformer (LVDT), has a very high deflection measurement accuracy of about 0.01 mm. However, the system requires additional supports at the bridge bottom to support such sensors, while most bridge bottoms are unsuitable for constructing such support systems, especially for long-term monitoring, due to river and traffic factors [12,13]. The global positioning system (GPS) method is convenient to install on bridges and is applied on many large-scale bridges. However, the deflection measurement accuracy is insufficient for short- or medium-span bridges, because its static deflection measurement accuracy is only about 20 mm, whereas the large deflection of simply supported girder bridges under vehicle loads is often several millimeters [14,15,16]. Furthermore, GPS is limited partly by multi-path and cycle slips, low frequency of data, as well as the need for good satellite coverage. Acceleration sensors are convenient to install at a relatively low cost, but they are not a good choice for monitoring bridge deflection. One reason is that they cannot be easily used for static deflection measurement when the main parts of bridge deflection under traffic loads are from the static loading of the vehicle dead-weight. Another reason is that the deflection assessment accuracy is easily influenced by bridge vibration and cumulative amplification error [17]. Image methods [18,19], such as digital image correlation (DIC) [20], can achieve millimeter-level accuracy or better in experimental environments, and is considered an expectable method in bridge deflection measurement. However, the accuracy depends largely on the high requirements of pixels and complicated systems, which makes the deflection monitoring system expensive.
Owing to the limitations of the aforementioned methods, strain sensor-based deflection measurement methods with low cost and easy application have been developed [21,22]. In these methods, bridge deflection is calculated using the strain–deflection relationship. However, to acquire accurate measurements, many strain sensors are frequently required, resulting in markedly increased deflection testing costs. In recent years, distributed long-gauge strain sensing technology has been developed to improve deflection measurement systems [23,24,25,26]. However, long-gauge strain sensors are still needed to cover most of the beam length. Therefore, the systems are still complicated and expensive. Besides the sensors, another factor, namely the strain–deflection relationship, will determine the performance of these indirect methods. The strain–deflection relationship is usually determined by the bridge structure mode. If the difference between the mode and actual structure is small, deflection assessment accuracy can be good. If the difference is large, the method may fail. Finally, factors such as temperature, loading, and damage, causing strain change, may also influence the performance of the indirect methods based on strain input. Therefore, the methods should be further studied for actual application.
This study presents a novel method for the deflection monitoring of simply supported girder bridges. The proposed method requires only a single long-gauge strain sensor to be installed at the location where deflection measurement is needed. The long-gauge strain sensor enables the collection of the temporal response of strain when a vehicle crosses the bridge. The temporal response of strain results can then be converted into the spatial distribution of strain along the bridge length. The converted strain distribution enables the easy calculation of deflection based on material and structural mechanics. Finally, the proposed method’s accuracy is validated using finite element (FE) simulations and on-site experiments.

2. Theory of Deflection Calculation Based on Strain Measurement Input

2.1. Deflection Calculation Based on Distributed Strain Measurement Input

In material mechanics, the strain ε(x) can be calculated from the known section’s bending moment M(x):
ε ( x ) = M ( x ) y E I z
where y is the distance from the strain measurement location to the neutral axis, E is the elastic modulus, and Iz is the moment of inertia. Deflection can then be calculated using the strain distribution measurement along the beam based on classical structural mechanics:
Δ = ε ( x ) M _ _ _ y d s
where M ¯ is the virtual bending moment obtained by implementing a one-unit load at the beam location that requires deflection measurement. Therefore, beam deflection can be calculated when the distributed strains of different sections ε(x) are known.

2.2. Assessment Method of Deflection Based on Strain Measurements at One Location under TrafficLoad

2.2.1. Displacement Reciprocal Theorem

The displacement reciprocal theorem is a universal theorem that explains the force–displacement relationship of a beam under Hooke’s law and small deformation. As shown in Figure 1, this theorem is as follows:
δ 12 = δ 21
If location 2 is further fixed in the beam middle and location 1 is moved, Equation (3) remains true. δ12 is the deflection at location 1 when the load F lies at the middle location, and δ21 is the deflection at the middle location when F moves to location 1. Therefore, the deflection at the beam middle when the moving load F acts on x equals that at x when the static load F acts on the middle location.
Furthermore, the deflection distribution along the beam length when the static load F acts at the middle location is equivalent to the measured deflection at the beam middle when the load F moves across the beam. Because curvature is the second-order integral of deflection, the law also applies to it. In other words, the curvature distribution along the beam length when the static load F acts at the middle location is equivalent to the measured curvature at the beam middle when the load F moves across the beam. This law is true for strain since it can be directly derived from the curvature. Therefore, this study further assumes that the strain distribution when traffic loads act at the beam middle is equivalent to the strain response when traffic loads move across the beam.

2.2.2. Static Strain Distribution Obtained with One Strain Sensor under Traffic Loads

(1)
Uniaxial vehicle load
When a vehicle wheelbase is markedly shorter than the bridge span, the vehicle load can be simplified into a uniaxial load. As shown in Figure 2a, the bending moment M(t) at the simply supported beam middle can be expressed as:
M ( t ) = F v t 2 0 t < L 2 v F L v t 2 L 2 v t L v
where L is the beam’s total length; t is the time; F is the vehicle load; and v is the vehicle velocity. Considering the vehicle location x(t) = vt, M(t) can be converted into an equivalent bending moment distribution at the middle location M(x(t)):
M ( x ( t ) ) = F x ( t ) 2 0 x ( t ) < L 2 F L x ( t ) 2 L 2 x ( t ) L
As shown in Figure 2b, when the uniaxial load is applied at the middle location, the bending moment distribution under the static load of the whole beam M(x) can be expressed as:
M ( x ) = F x 2 0 x < L 2 F L x 2 L 2 x L
Comparing Equations (5) and (6) shows that M(x) and M(x(t)) have the same form when x(t) and x mean the same location. The static strain distribution ε(x) can be obtained by Equation (7):
ε ( x ) = M ( x ) y E I z = M ( x ( t ) ) y E I z = ε ( x ( t ) )
where ε(x(t)) is the equivalent strain distribution at the middle location. Therefore, the equivalent strain distribution can be substituted for the static strain distribution in Equation (2) to calculate deflection.
(2)
Biaxial vehicle load
Most short- and medium-sized bridges cannot accommodate the simplification of a vehicle load into a uniaxial load. The whole process for a vehicle load passing over a simply supported beam can be divided into five stages (Figure 3), where F1 is the front axle load, F2 is the rear axle load, and m is the wheelbase.
Similarly, the bending moment M(t) of the middle location of a simply supported beam can be expressed as:
M ( t ) = F 1 2 v t 0 t < m v F 1 2 v t + F 2 2 ( v t m ) m v t < L 2 v F 1 2 ( L v t ) + F 2 2 ( v t m ) L 2 v x ( t ) < L + 2 m 2 v F 1 2 ( L v t ) + F 2 2 ( L v t + m ) L + 2 m 2 v x ( t ) < L v F 2 2 ( L v t + m ) L v x ( t ) L + m v
Let x(t) be the location of the combined force F1 + F2 (Figure 4) of the front axle F1 and rear axle F2 at time t. For brevity, let F 1 F 1 + F 2 = α , F 2 F 1 + F 2 = β . Then, x(t) can be represented as:
x ( t ) = v t F 2 F 1 + F 2 m = v t β m
Substituting Equation (9) into Equation (8) yields an equivalent bending moment distribution at the middle location M(x(t)) as follows:
M ( x ( t ) ) = F 1 2 ( x ( t ) + β m ) β m x ( t ) < α m ( F 1 + F 2 ) 2 x ( t ) α m x ( t ) < L 2 β m ( F 2 F 1 ) x ( t ) 2 α F 2 m + F 1 L 2 L 2 β m x ( t ) < L 2 + α m ( F 1 + F 2 ) 2 ( L x ( t ) ) L 2 + α m x ( t ) < L β m F 2 2 ( L x ( t ) + α m ) L β m x ( t ) L + α m
For comparison, the locations of the front and rear axles are placed in reverse order, and the combined force is located at the middle location (Figure 5). The bending moment distribution of the whole beam under static biaxial vehicle load M(x) can be expressed as follows:
M ( x ) = ( F 1 + F 2 ) 2 x 0 x < L 2 β m ( F 2 F 1 ) x 2 α F 2 m + F 1 L 2 L 2 β m x < L 2 + α m ( F 1 + F 2 ) 2 ( L x ) L 2 + α m x < L
Comparing Equations (10) and (11) reveals that M(x(t)) equals M(x) when x = x(t) and x∈[αm, L − βm]. If the beam stiffness EI and height of the neutral axis y remain unchanged, then ε(x(t)) = ε(x). Therefore, the equivalent strain distribution ε(x(t)) can be substituted for the static strain distribution ε(x) into Equation (2) to calculate deflection directly. When x∈[−βm, αm] and x∈[Lβm, L + αm], a modified method is needed to improve accuracy.
The deflection Δ1 calculated from the static strain distribution when x∈[0, αm] and x∈[Lβm, L] and the virtual bending moment M ¯ ( x ) are as follows:
Δ 1 = 0 α m ε ( x ) M ¯ ( x ) y d x + L β m L ε ( x ) M ¯ ( x ) y d x
M ¯ ( x ) = 1 2 x 0 x < L 2 1 2 ( L x ) L 2 x L
The deflection Δ2 calculated from the equivalent strain distribution when x ∈ [−βm, αm] and x ∈ [Lβm, L + αm] is:
Δ 2 = β m α m ε ( x ( t ) ) M ¯ ( x ( t ) ) y d x ( t ) + L β m L + α m ε ( x ( t ) ) M ¯ ( x ( t ) ) y d x ( t ) = 0 m ε ( x ( t ) β m ) M ¯ ( x ( t ) β m ) y d x ( t ) + L m L ε ( x ( t ) + α m ) M ¯ ( x ( t ) + α m ) y d x ( t )
where M ¯ ( x ) is the modified virtual bending moment. Let Δ1 = Δ2, then the following is obtained:
0 m ε ( x ( t ) β m ) M ¯ ( x ( t ) β m ) y d x ( t ) = 0 α m ε ( x ) M ¯ ( x ) y d x L m L ε ( x ( t ) + α m ) M ¯ ( x ( t ) + α m ) y d x ( t ) = L β m L ε ( x ) M ¯ ( x ) y d x
When x ∈ [0, αm] and x = x(t), then
ε ( x ) ε ( x ( t ) β m ) = M ( x ) M ( x ( t ) β m ) = ( F 1 + F 2 ) 2 x F 1 2 x ( t ) = F 1 + F 2 F 1 = 1 α
When x ∈ [L − βm, L] and x = x(t), then
ε ( x ) ε ( x ( t ) β m ) = M ( x ) M ( x ( t ) β m ) = ( F 1 + F 2 ) 2 x F 1 2 x ( t ) = F 1 + F 2 F 1 = 1 α
M ¯ ( x ) can then be further expressed as:
M ¯ ( x ) = M ¯ ( x + β m ) 1 α β m x < ( α β ) m 0 ( α β ) m x < α m M ¯ ( x ) α m x < L β m 0 L β m x < L + ( α β ) m M ¯ ( x α m ) 1 β L + ( α β ) m x < L + α m
Therefore, if the virtual bending moment M ¯ ( x ) is represented as above, deflection calculation errors can be eliminated.
(3)
Multi-axle vehicle load
The proposed method also applies to vehicles with at least three axles. Here, the three-axle vehicle is taken as an example to explain the calculation process. As shown in Figure 6, F1, F2, and F3 are the loads for the 3 axles, respectively; m 1 and m 2 are the wheelbases; p is the combined force location and p = m 1 F 2 + ( m 1 + m 2 ) F 3 F 1 + F 2 + F 3 .
The equivalent bending moment distribution at the middle location M(x(t)) can further be expressed as:
M ( x ( t ) ) = F 1 2 ( x ( t ) + p ) p < x ( t ) p + m 1 ( F 1 + F 2 ) ( x ( t ) + p ) F 2 m 1 2 p + m 1 < x ( t ) p + m 1 + m 2 F 1 + F 2 + F 3 2 x ( t ) p + m 1 + m 2 < x ( t ) L 2 m 1 m 2 + p ( F 1 + F 2 F 3 ) 2 x ( t ) + F 3 ( L 2 m 1 m 2 + p ) L 2 m 1 m 2 + p < x ( t ) L 2 m 1 + p F 1 + F 2 + F 3 2 ( L x ( t ) ) F 1 ( L 2 x ( t ) + p ) L 2 m 1 + p < x ( t ) L 2 + p F 1 + F 2 + F 3 2 ( L x ( t ) ) L 2 + p < x ( t ) L p ( F 2 + F 3 ) ( L + m 1 + m 2 p x ( t ) ) F 2 m 2 2 L p < x ( t ) L p + m 1 F 3 ( L + m 1 + m 2 p x ( t ) ) 2 L p + m 1 < x ( t ) L + m 1 + m 2 p
The bending moment distribution under static biaxial vehicle load M(x) can be expressed as follows:
M ( x ) = F 1 + F 2 + F 3 2 x 0 < x L 2 m 1 m 2 + p ( F 1 + F 2 F 3 ) 2 x + F 3 ( L 2 m 1 m 2 + p ) L 2 m 1 m 2 + p < x L 2 m 1 + p F 1 + F 2 + F 3 2 ( L x ) F 1 ( L 2 x + p ) L 2 m 1 + p < x L 2 + p F 1 + F 2 + F 3 2 ( L x ) L 2 + p < x L
M(x(t)) still equals M(x) when x = x(t) and x ∈ [m1 + m2p, Lp]. Aiming to eliminate errors, M ¯ ( x ) can also be expressed as:
M ¯ ( x ) = M ¯ ( x + p ) 1 η p x < m 1 + m 2 2 p 0 m 1 + m 2 2 p x < m 1 + m 2 p M ¯ ( x ) m 1 + m 2 p x < L p 0 L p x < L + m 1 + m 2 2 p M ¯ ( x m 1 m 2 + p ) 1 γ L + m 1 + m 2 2 p x < L + m 1 p M ¯ ( x m 1 m 2 + p ) 1 λ L + m 1 p x < L + m 1 + m 2 p
The parameters are expressed as:
1 η = F 1 + F 2 + F 3 F 1 1 γ = F 1 + F 2 + F 3 F 2 + F 3 1 λ = F 1 + F 2 + F 3 F 3

3. Strain Measurement Based on Optical Fibre Sensors

3.1. Strain Measurement Principle

Owing to their high precision and serializability, fibre Bragg grating (FBG) strain sensors are often used in engineering to measure the strain of structures. As shown in Figure 7, an FBG sensor comprises a cladding layer, coating layer, fibre core, and grating region. When input light passes through the fibre grating, the grating will reflect back some light satisfying the Bragg condition, whereas the remaining transmitted light will continue to propagate along the fibre core. If some strain ε or temperature T change happens upon the grating region, the distance of grating Λ will change. Then, the central wavelength of the reflected light will change. A linear relationship between wavelength variation and strain is given as follows:
λ B = α ε ε + α T T
where λ B is the central wavelength change in reflected light; α ε is the strain sensitivity coefficient, which is 1.2 pm/με generally; αT is the temperature sensitivity coefficient.

3.2. Long-Gauge Optical Fibre Sensor

The size of a bare FBG sensor is approximately 2 cm and may not be ideal for monitoring concrete structures, due to the nonuniform material characteristics of concrete. Moreover, a steel material is often used to package the bare FBG sensor since it is brittle and difficult to use without treatment. However, the bonding surface between the inner optical fibre and outer steel is not reliable over the long term. Therefore, the authors’ research team has proposed a new long-gauge FBG sensor using a fibre-reinforced polymer (FRP) package, as shown in Figure 8, with a gauge length of up to 1 m. Owing to strain uniformity along the optical fibre within the sensing range, the strain will reflect the average deformation of the monitored structure throughout this range, thereby resulting in a significant reduction in the negative influence on strain measurement caused by material inhomogeneity and local damage, such as concrete cracks. Furthermore, the long-term performance of FRP and the bonding surface between optical and reinforced fibres are excellent.

3.3. Strain Sensing Performance of the Sensor

Because the long-gauge FBG sensor is cross-sensitive to temperature and strain, the sensors were subjected to tensile and temperature tests [25]. The test results shown in Figure 9 reveal that the wavelength–strain relationship is highly linear. The sensor has good repeatability since the correlation of the average linear fitting exceeds 0.999. The strain sensing coefficient is 1.18 pm/με, which is very close to the theoretical value of 1.2 pm/με. The linearity and repeatability of the sensor’s wavelength and temperature are also good, with the correlation of the average linear fitting exceeding 0.99.

4. Study on the FE Bridge Model

4.1. Introduction to the Bridge Model

One typical highway bridge is selected as an example to assess the performance of the proposed method. The bridge comprises 12 prefabricated hollow prestressed reinforced concrete (PRC) girders with a span of 20 m. The girders are combined with each other with one hinge joint made of reinforced concrete (RC). The dimensions are as shown in Figure 10. Then the finite element (FE) model of the simply supported girder bridge is established using Midas FEA (Figure 11). In the FE model, the girders, and hinge joints are simulated with solid elements. The constraints at both ends are simply supported by imposing three degrees of freedom on the left (Dx, Dy, and Dz) and two degrees of freedom on the right (Dx and Dz). In this paper, the steel rebar and prestressed rebar are not included in the model for simplification, as the elastic model is sufficient to verify the proposed method. The detailed information of the material properties of FE models are listed in Table 1.
In this study, the loads of one typical truck are placed on the surface of the middle girder (the sixth one). The front axle load of the vehicle is 25 kN, whereas the rear axle load is 100 kN. The wheelbase is 4 m. The driving speed is 30 km/h. As shown in Figure 12, static and vehicle moving loads are applied respectively, in which the dotted lines represent the combined force.

4.2. Extraction and Processing of Strain Data

The girder is divided into 20 measurement units, each of which is 1 m long. The strain measurement by long-gauge FBG sensors is simulated to extract the average strain of each unit, called the macro-strain. The simulated sensor distribution is shown in Figure 13. Here, there is no actual sensor model in the FE model, while the strain is just extracted at the position at which the designed sensor lies. For each sensor, the average strain within the gauge length is obtained from the model directly, which is taken as the measured macro-strain.
Apart from truck weight-induced strain, dynamic strain contains some additional vibration response, which inevitably causes interference and error. Therefore, in actual signal processing, signal denoising is required to obtain an accurate quasi-static strain response.
Traditional denoising methods, such as the fast Fourier transform (FFT), cannot well distinguish the useful signal within a high frequency from the high frequency part caused by noise, especially for the case of non-stationary signals whose frequency changes with time. Luckily, the improvement of a wavelet denoising method can be used to distinguish the useful signal. In the wavelet denoising method, the traditional basis function with an infinite length is replaced by a finite length attenuation wavelet basis (Figure 14). The change makes it easy to know not only the frequency component of the signal, but also its specific location in the time domain.
The basic idea of wavelet denoising is to remove noise components and leave key information by decomposition and selective reconstruction. The specific expression of wavelet denoising is as follows.
W T ( a , τ ) = 1 a f ( t ) Ψ ( t τ a ) d t
where τ is time shift, a is the scaling factor, Ψ(t) is the wavelet basis function.
The basic steps of wavelet denoising are as follows:
  • Wavelet decomposition. Generally, a wavelet basis type and the layer N of wavelet decomposition are selected first, and then N layer decomposition of the noisy signal is performed to obtain the wavelet coefficient of each layer. The wavelet basis function used in this paper is Sym8, which has already been used in many application scenarios in signal processing and data compression. Sym8 can decompose the data signal into multiple sub-signals, reducing the data noise while retaining the important features of the signal. It can retain the important features of the signal as much as possible and remove the noise interference. Here N is set as 4.
  • Threshold processing. The wavelet coefficients from the first layer to the N layer are processed with appropriate thresholds, and the noise is removed with its de-correlation (Figure 15). The main principle is to set a threshold λ. When the wavelet coefficient is greater than λ, it retains its original value. When the wavelet coefficient is less than λ, it is directly set to 0. In this way, the original features are well preserved and the noise removal signal is smoother.
  • Wavelet reconstruction. The processed signal is reconstructed and the denoised signal is obtained. In Figure 16, the signal S(t) is decomposed by four-layer wavelet, in which WCD1, WCD2, WCD3, and WCD4 are the high-frequency wavelet coefficients of each layer wavelet decomposition, while WCA1, WCA2, WCA3, and WCA4 are the low-frequency wavelet coefficients. Here, noise is mainly contained in the low-frequency wavelet coefficient, so it is considered that WCA1, WCA2, WCA3, and WCA4 are less than λ, and then can be set as 0. Finally, the denoised signal is obtained.
In this study, the signal processing method of wavelet transform is applied to denoise the dynamic strain. Figure 17 depicts the typical results of signal denoising.

4.3. Results and Analysis

Owing to the symmetry of the studied bridge, half of the girders are selected for analysis. The denoised strain responses are transformed into the equivalent strain distribution based on the method described in Section 2, which is then compared with the static strain distribution directly obtained from the FE simulation (Figure 18). The sampling frequency is set as 200 Hz. The findings show that the proposed method is effective despite some minor discrepancies, particularly at the two ends, which occur due to the partial and full exertion of moving biaxial and static loads on the girder and the inability of the denoising process to delete all dynamic responses.
The deflection at the girder midpoint can then be calculated by inputting the strain results into Equation (2). Table 2 lists the deflection calculation results; notably, the actual deflection is derived directly from the FE simulation. Most deflection errors are below 5%, indicating that the proposed method has a high degree of accuracy. Moreover, the accuracy based on the input of the equivalent strain distribution slightly exceeds that based on the input of the static strain distribution. However, the highest error occurs at the side girder for both strain input types, but the reason for this is unclear. This may be caused by the FE simulation itself, as the side girder has an unrestrained side, making the FE calculation more complicated.

4.4. Parameter Influence Analysis

For actual bridges, some parameters can influence the proposed method’s effectiveness. Therefore, parameters such as vehicle type, vehicle speed, and structural damage are investigated in this study.
(1)
Vehicle type
Based on different axle weights, wheelbases, and number of axles, vehicle types can be roughly divided into the following categories: biaxial van, biaxial truck, three-axle truck, and multi-axle trailer. Figure 19 shows the axle weights and wheelbases of the different vehicle types.
The vehicle parameters are incorporated into the FE model. The static and equivalent strain distributions can then be obtained, as shown in Figure 20. Compared with the other two types of vehicles, the degree of coincidence between the two types of strain results produced by a biaxial van and a biaxial truck is greater in the figures. At the two ends of the figures, notably for the multi-axle trailer, there are obvious variances due to the partial and total exertion of moving axle and static loads on the girder. This situation is more obvious when the wheelbases are long. Due to their longer wheelbases, multi-axle trailers have the highest error.
Similarly, deflection is calculated using the two types of strain results (Table 3), with only girder 6#’s results included. The results indicate that vehicle type influences the proposed method’s accuracy. However, the error caused by vehicle type differences can be controlled since the highest error is approximately 5%.
(2)
Vehicle speed
Four typical speeds are included in the FE model to investigate how vehicle speed affects deflection measurement accuracy: 30, 60, 90, and 120 km/h. The difference between both strain types increases as vehicle speed increases, as shown in Figure 21. The main reason is that a higher vehicle speed will cause the bridge to vibrate more severely. In addition, the denoising process can only delete a portion of the vibration strain response because the strain curves are no longer smooth and fluctuate slightly after denoising.
Table 4 shows the calculated deflection results at different vehicle speeds. The results show that the calculated deflection error also increases as vehicle speed increases. However, the error caused by differences in vehicle speed can also be controlled since the highest error is 5.5% at 120 km/h.
(3)
Structural damage
A decrease in local stiffness induced by structural deterioration, such as damage to hinge joints and concrete cracks, will impact strain. Consequently, these damage types are used as examples to examine the influence of structural damage on the proposed deflection measurement method.
 (1)
Damage between the hinge joints
Partial hinge joint damage is typical damage of a simply supported girder bridge in China, which greatly decreases the entire bridge stiffness. In the FE model, hinge joint damage is simulated by decreasing the elastic modulus. The degree of hinge joint damage is calculated using the length of damaged hinge joints. The damage factor D is then defined as follows:
D = l L
where l is the length of the damaged hinge joints, and L is the total length of the hinge joints. Notably, hinge joints are usually destroyed from the middle part outwards, due to the larger load response there. Therefore, damage to the hinge joint is assumed to start symmetrically from the middle unit to the two ends (Figure 22). The damage factor D of the 5# hinge joint is set from 0.1 to 1.0.
As illustrated in Figure 23, the development of hinge joint damage has a substantial effect on strain results. As damage accumulates, the disparity between the two strain distribution types widens.
Table 5 shows the deflection results for hinge joint damage developing from the middle part. When the damage factor D is less than 0.5, which means that the damage only occurs at the small middle part, the deflection error calculated by equivalent strain distribution is the largest, reaching a peak value of about 18%. However, if the damage factor D exceeds 0.5, the calculation error decreases significantly as the damage develops, because the local stiffness of a bridge differs along the girder due to local hinge joint damage. In the proposed method, stiffness is assumed to be uniform or known along the girder. If the changed stiffness distribution is disregarded in the proposed method when converting the strain response to the equivalent strain distribution, an obvious error will occur as a result. Therefore, if damage to hinge joints has already occurred at the middle parts, especially where the strain sensor is installed, stiffness distribution should be modified to improve deflection calculation accuracy.
Due to a defective bearing system, the hinge joint occasionally suffers damage at the girder end. Therefore, some simulations are conducted to address such a case (Figure 24). The damage factor D is set from 0.05 to 0.45. The typical strain results are shown in Figure 25. The difference between the two strain distribution types also tends to widen as the damage develops, but not as clearly as identified above. This difference is quite small, indicating that the damaged hinge joint at the girder’s edge has little effect on the proposed method. This result is validated by the deflection calculation results (Table 6), which show the highest error of calculated deflection as about 4%. The primary reason is because the strain response is measured at the girder’s middle, which is minimally affected by local damage at the girder’s edge. Therefore, assuming that damage to the hinge joint only occurs on one side, the proposed method for measuring deflection is promising.
 (2)
Concrete cracks
Two types of concrete cracks at the bottom of actual bridge girders are longitudinal and transverse. Prestress-induced cracks are longitudinal, whereas bending moment-induced cracks are transverse. As seen in Figure 26, the FE model includes a longitudinal crack over the whole girder length and a transverse crack across the entire girder width. The width of the two cracks is 0.4 mm, whereas the depth is 50 mm. Crack simulation is achieved by removing entities of corresponding shapes.
The strain results are shown in Figure 27, and the deflection results are listed in Table 7. Notably, concrete cracks have a negligible effect on the proposed method since the equivalent strain is close to the static strain and the calculated deflection error is below 3%. In general, longitudinal cracks have little impact on a girder’s bending stiffness, but transverse cracks decrease bending stiffness. Fortunately, due to the long-gauge strain sensor, the transverse crack’s effect on the average bending stiffness within the long-sensing gauge will be minimal. Therefore, when transverse cracks are few, their effects on the proposed method’s accuracy can be ignored.

5. On-Site Experimental Study

5.1. Experimental Description

In this study, the experimental bridge is a standard simply supported girder bridge with a span of 20 m on China’s National Highway G204. The bridge is separated into two independent parts for two traffic directions: south and north. One side of the bridge is made of 12 precast girders and is separated into a pavement, a heavy lane, and a fast lane (Figure 28). Strain sensors are installed on the 7# girder’s bottom since this location experiences heavy traffic loads.
The 7# girder is divided into 20 monitoring units, each of which is 1 m long. The gauge length of the long-gauge FBG sensors (shown in Figure 8) used in the experiments is also 1 m, and then each sensor can cover each girder unit. Despite the difficulty in installing sensors due to space limits, 16 sensors were used in the actual experiments (Figure 29). The sensors were prepared in two groups, each of which had 8 FBG sensors (Figure 29a). For comparison, the strains of units without sensors were interpolated based on the strains detected by nearby sensors. In the experiments, the FBG sensors were adhered to the selected girder’s bottom surface with epoxy resin (Figure 30a). Furthermore, some displacement meters were installed with scaffolds to measure the actual deflection (Figure 30b).
Static and moving vehicle loads were applied in the experiments (Figure 31). A truck full of sand, about 25 t, was used to implement the loading tests. The wheelbase between the two axles was 4.2 m. When the static load was applied, the midpoint of the two axles was placed at the bridge midpoint. When the moving vehicle load was applied, the truck passed over the bridge at 30 and 60 km/h. Planned speeds of 90 and 120 km/h were also established, although these were difficult to achieve owing to traffic limits. The sampling frequency was 200 Hz.

5.2. Experimental Results and Analysis

Figure 32 shows the typical strain results obtained before and after denoising. Certain signal noise and dynamic signals can be observed via the denoising process. Figure 33 compares the strain distribution under static loading tests to the comparable strain distribution derived from the strain response. Speed has some effect on the accuracy of strain transformation since the difference between the two strain distribution types widens as driving speed increases. This effect occurs because increasing the speed causes the bridge to vibrate more and the denoising process can delete only a part of the dynamic strain response.
Table 8 lists the results of the deflection at the girder’s midpoint calculated using Equation (2) and the strain distribution. The proposed method is deemed to have satisfactory deflection measurement accuracy because the largest relative error is roughly 8%. Meanwhile, the largest absolute error of the deflection measurement is 0.05 mm, which is small enough. Overall, these findings suggest that the proposed method is an effective technique for measuring deflection, notably for short- or medium-span concrete bridges, where deflection is often minimal (several millimetres or smaller than one millimetre).

6. Conclusions and Remarks

In this study, a novel method for quick deflection measurement of a simply supported girder bridge based on strain response under traffic loads was developed. The strain conversion theory was first constructed for the temporal strain response and spatial distribution of strain. The accuracy of the proposed deflection measurement method was then validated using FE simulations and on-site experiments of a typical bridge. Consequently, the following are the key conclusions:
(1)
The displacement reciprocal theorem enabled the development of a method for converting the temporal strain response under traffic loads to the spatial distribution of strain under the same static loads. Meanwhile, modified methods for decreasing the error caused by vehicle type differences were established.
(2)
The deflection measurement method relied largely on strain distribution input, which was measured using the proposed macro-strain sensor, a long-gauge FBG sensor. The high accuracy of the proposed method was validated by FE simulation results, which indicated a relative error of about 5%. However, when key parameters such as vehicle type, vehicle speed, and structural damage were considered, damage to the hinge joint in the middle location significantly affected the proposed method’s accuracy, as the deflection error could exceed 10%.
(3)
The results of on-site experiments validated the proposed method, as the maximum relative error of deflection measurement was about 8% and the absolute error was below 0.05 mm.
Although the proposed method is effective for measuring deflection, especially for short- and medium-span concrete bridges where deflection is often minimal (several millimetres or smaller than one millimetre), due to time constraints, future studies are required to expand its application. First, more on-site tests should be conducted and completed using different vehicle types to further prove the proposed method’s accuracy. Second, long-term performance, such as repeatability of strain measurement, sensor life, and the sensor–girder bonding surface should be given greater consideration. Third, material degradation should also be considered in the equations when the method is used for long-term monitoring.
Considering cost and efficiency, the proposed method is a good choice for measuring the deflection of short- and medium-span concrete bridges, although further research must be conducted in the future. In addition, certain benefits will be realized given the superior long-term performance of optical fibre sensing technology; hence, the proposed method has a promising future in practical applications.

Author Contributions

Conceptualization, Y.T. and J.C.; Data curation, Y.T. and J.C.; Formal analysis, J.C. and B.Z.; Funding acquisition, Y.T.; Investigation, J.C., B.Z. and W.T.; Methodology, Y.T.; Validation, Y.T. and J.C.; Writing—original draft, J.C.; Writing—review and editing, Y.T., B.Z. and W.T. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the financial support by a grant from the Engineering Research Center of Intelligent Construction and Industrialization, CAAC (Grant No. MHJGKFKT-03) and another grant from the National Key R&D Program of China (Grant No. 2021YFB2601200).

Data Availability Statement

All data included in this study are available upon request by contact with the corresponding author. The data are not publicly available due to privacy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, D.H.; Zhou, H.; Yi, T.H.; Li, H.N.; Bai, H.F. Joint deterioration detection based on field-identified lateral deflection influence lines for adjacent box girder bridges. Struct. Control Health Monit. 2022, 29, e3053. [Google Scholar] [CrossRef]
  2. Cusson, D.; Lounis, Z.; Daigle, L. Durability monitoring for improved service life predictions of concrete bridge decks in corrosive environments. Comput. Aided Civ. Infrastruct. Eng. 2011, 26, 524–541. [Google Scholar] [CrossRef]
  3. Wen, X.L.; Wei, L.; Dan, D.H.; Liu, G.M. Study on a measurement index of transverse collaborative working performance of prefabricated girder bridges. Adv. Struct. Eng. 2017, 20, 1879–1890. [Google Scholar] [CrossRef]
  4. Zhang, L.; Feng, D.M.; Wu, G. Simultaneous identification of bridge damage and vehicle parameters based on bridge strain responses. Struct. Control Health Monit. 2022, 29, e2945. [Google Scholar] [CrossRef]
  5. Daniel, M.; Abdollah, M.; Eugene, O.B. Bridge health monitoring using deflection measurements under random traffic. Struct. Control Health Monit. 2020, 27, e2593. [Google Scholar]
  6. Zhu Ge, S.; Xu, X.P.; Zhong, L.J.; Gan, S.W.; Lin, B.; Yang, X.; Zhang, X.H. Noncontact deflection measurement for bridge through a multi-UAVs system. Comput. Aided Civ. Infrastruct. Eng. 2021, 37, 746–761. [Google Scholar] [CrossRef]
  7. Jacek, K.; Wojciech, A.; Damian, B. Comparison of non-destructive techniques for technological bridge deflection testing. Materials 2020, 13, e1908. [Google Scholar]
  8. Oleksandr, P.M.; Oleksandr, M.S.; Yaroslav, L.I.; Petro, S.K. Multilaser spot tracking technology for bridge structure displacement measuring. Struct. Control Health Monit. 2021, 28, e2675. [Google Scholar]
  9. Erdoğan, H.; Gülal, E. Ambient Vibration Measurements of the Bosphorus Suspension Bridge by Total Station and GPS. Struct. Control Health Monit. 2013, 37, 16–23. [Google Scholar] [CrossRef]
  10. Pan, X.D.; Zhao, L. Laboratory experiment study on real-time monitoring and safety early warning system of structure settlement and deformation. Appl. Mech. Mater. 2015, 775, 264–267. [Google Scholar] [CrossRef]
  11. Lee, Z.K.; Bonopera, M.; Hsu, C.C.; Lee, B.H.; Yeh, F.Y. Long-term deflection monitoring of a box girder bridge with an optical-fibre, liquid-level system. Structures 2022, 44, 904–919. [Google Scholar] [CrossRef]
  12. Santhosh, K.V.; Roy, B.K. Online implementation of an adaptive calibration technique for displacement measurement using LVDT. Appl. Soft Comput. 2017, 53, 19–26. [Google Scholar]
  13. Sarthak, J.; Shrikant, M.H. Linear variable differential transducer (LVDT) & its applications in civil engineering. Int. J. Transp. Eng. Technol. 2017, 3, 62–66. [Google Scholar]
  14. Xu, Y.; James, M.W.B.; David, H.; Koo, K.Y. Long-span bridges: Enhanced data fusion of GPS displacement and deck accelerations. Eng. Struct. 2017, 147, 639–651. [Google Scholar] [CrossRef]
  15. GEsteban, V.B.; JRamon, G.C.; Rick, B.; Michel, G.A.; Ivan, E.G.C. Structural evaluation of dynamic and semi-static displacements of the Juarez Bridge using GPS technology. Measurement 2017, 110, 146–153. [Google Scholar]
  16. Hussein, A.M.; Panos, A.P.; Craig, M.H.; Gethin, W.R.; Lukasz, B. Monitoring the response of Severn Suspension Bridge in the United Kingdom using multi-GNSS measurements. Struct. Control Health Monit. 2021, 28, e2830. [Google Scholar]
  17. Meng, X.; Dodson, A.H.; Roberts, G.W. Detecting bridge dynamics with GPS and triaxial accelerometers. Eng. Struct. 2007, 29, 3178–3184. [Google Scholar] [CrossRef]
  18. Rujika, T.; Punchet, T.; Said, E. CCTV camera array for the displacement and strain measurement of a beam specimen in a laboratory. Buildings 2022, 12, 1778. [Google Scholar]
  19. Francisco, B.; Susana, A.; Pedro, J.S.; António, C.; Paulo, J.T.; Pedro, M.G.P.M.; Ranzal, D.; Cardoso, N.; Fernandes, N.; Fernandes, R.; et al. Displacement monitoring of a pedestrian bridge using 3D digital image correlation. Procedia Struct. Integr. 2022, 37, 880–887. [Google Scholar]
  20. Marie, A.; Olivier, P.; Noémie, P.; Emile, R.; Pierre, V. In situ DIC method to determine stress state in reinforced concrete structures. Measurement 2023, 210, 112483. [Google Scholar]
  21. Sun, L.; Li, Y.X.; Zhang, W. Experimental study on continuous bridge-deflection estimation through inclination and strain. J. Bridge Eng. 2020, 25, e04020020. [Google Scholar] [CrossRef]
  22. Hong, W.; Qin, Z.; Lv, K.; Fang, X. An indirect method for monitoring dynamic deflection of beam-like structures based on strain responses. Appl. Sci. 2018, 8, e811. [Google Scholar] [CrossRef]
  23. Feng, X.; Sun, C.S.; Zhang, X.T.; Ansari, F. Determination of the coefficient of thermal expansion with embedded long-gauge fibre optic sensors. Meas. Sci. Technol. 2010, 21, e065302. [Google Scholar] [CrossRef]
  24. Tang, Y.S.; Wu, Z.S. Distributed long-gauge optical fibre sensors based self-sensing FRP bar for concrete structure. Sensors 2016, 16, 286–303. [Google Scholar] [CrossRef]
  25. Tang, Y.S.; Jiang, T.F.; Wan, Y. Structural monitoring method for RC column with distributed self-sensing BFRP bars. Case Stud. Constr. Mater. 2022, 17, e01616. [Google Scholar] [CrossRef]
  26. Tang, Y.S.; Cao, M.F.; Li, B.; Chen, X.H.; Wang, Z.Y. Horizontal Deformation Monitoring of Concrete Pile with FRP-Packaged Distributed Optical-Fibre Sensors. Buildings 2023, 13, 2454. [Google Scholar] [CrossRef]
Figure 1. Displacement reciprocal theorem: (a) first state; (b) second state.
Figure 1. Displacement reciprocal theorem: (a) first state; (b) second state.
Buildings 14 00070 g001
Figure 2. Uniaxial vehicle load: (a) moving load; (b) static load.
Figure 2. Uniaxial vehicle load: (a) moving load; (b) static load.
Buildings 14 00070 g002
Figure 3. Process of biaxial load passing over a bridge: (a) only the front axle acts on the beam; (b) the front and rear axles both act on the beam, and the front axle does not exceed the middle location; (c) the front and rear axles both act on the beam, but only the front axle exceeds the middle location; (d) the front and rear axles both act on the beam and exceed the middle location; (e) only the rear axle acts on the beam.
Figure 3. Process of biaxial load passing over a bridge: (a) only the front axle acts on the beam; (b) the front and rear axles both act on the beam, and the front axle does not exceed the middle location; (c) the front and rear axles both act on the beam, but only the front axle exceeds the middle location; (d) the front and rear axles both act on the beam and exceed the middle location; (e) only the rear axle acts on the beam.
Buildings 14 00070 g003
Figure 4. Combined force of the front axle F1 and rear axle F2.
Figure 4. Combined force of the front axle F1 and rear axle F2.
Buildings 14 00070 g004
Figure 5. Static biaxial vehicle load.
Figure 5. Static biaxial vehicle load.
Buildings 14 00070 g005
Figure 6. Multi-axle vehicle load: (a) moving load; (b) static load.
Figure 6. Multi-axle vehicle load: (a) moving load; (b) static load.
Buildings 14 00070 g006
Figure 7. Strain measurement principle of FBG: (a) light transmission within an FBG sensor; (b) input light spectrum; (c) reflected light spectrum; (d) transmission light spectrum.
Figure 7. Strain measurement principle of FBG: (a) light transmission within an FBG sensor; (b) input light spectrum; (c) reflected light spectrum; (d) transmission light spectrum.
Buildings 14 00070 g007
Figure 8. Long-gauge FBG sensor.
Figure 8. Long-gauge FBG sensor.
Buildings 14 00070 g008
Figure 9. Results of the tests on the long-gauge FBG sensor: (a) tensile test; (b) temperature test.
Figure 9. Results of the tests on the long-gauge FBG sensor: (a) tensile test; (b) temperature test.
Buildings 14 00070 g009
Figure 10. Bridge properties and cross sections.
Figure 10. Bridge properties and cross sections.
Buildings 14 00070 g010
Figure 11. FE bridge model by Midas FEA: (a) prefabricated hollow girders; (b) hinge joints.
Figure 11. FE bridge model by Midas FEA: (a) prefabricated hollow girders; (b) hinge joints.
Buildings 14 00070 g011
Figure 12. Load application: (a) vehicle moving load; (b) static load.
Figure 12. Load application: (a) vehicle moving load; (b) static load.
Buildings 14 00070 g012
Figure 13. Strain extraction with simulated sensors: (a) static strain; (b) dynamic strain.
Figure 13. Strain extraction with simulated sensors: (a) static strain; (b) dynamic strain.
Buildings 14 00070 g013
Figure 14. Basic function of FFT and wavelet denoising.
Figure 14. Basic function of FFT and wavelet denoising.
Buildings 14 00070 g014
Figure 15. Wavelet denoising basic steps.
Figure 15. Wavelet denoising basic steps.
Buildings 14 00070 g015
Figure 16. Wavelet reconstruction schematic.
Figure 16. Wavelet reconstruction schematic.
Buildings 14 00070 g016
Figure 17. Typical results of signal denoising.
Figure 17. Typical results of signal denoising.
Buildings 14 00070 g017
Figure 18. Comparison results between static and equivalent strain distributions: (a) 1#; (b) 2#; (c) 3#; (d) 4#;(e) 5#; (f) 6#.
Figure 18. Comparison results between static and equivalent strain distributions: (a) 1#; (b) 2#; (c) 3#; (d) 4#;(e) 5#; (f) 6#.
Buildings 14 00070 g018
Figure 19. Different vehicle types: (a) biaxial van; (b) biaxial truck; (c) three-axle truck; (d) multi-axle trailer.
Figure 19. Different vehicle types: (a) biaxial van; (b) biaxial truck; (c) three-axle truck; (d) multi-axle trailer.
Buildings 14 00070 g019
Figure 20. Strain results of different vehicle types: (a) biaxial van; (b) biaxial truck; (c) three-axle truck; (d) multi-axle trailer.
Figure 20. Strain results of different vehicle types: (a) biaxial van; (b) biaxial truck; (c) three-axle truck; (d) multi-axle trailer.
Buildings 14 00070 g020
Figure 21. Strain results at different speeds: (a) 30; (b) 60; (c) 90; (d) 120 km/h.
Figure 21. Strain results at different speeds: (a) 30; (b) 60; (c) 90; (d) 120 km/h.
Buildings 14 00070 g021
Figure 22. Damage to the hinge joint in the middle location: (a) overall; (b) detail.
Figure 22. Damage to the hinge joint in the middle location: (a) overall; (b) detail.
Buildings 14 00070 g022
Figure 23. Typical strain results for hinge joint damage developing from the middle part: (a) D = 0.2; (b) D = 0.8.
Figure 23. Typical strain results for hinge joint damage developing from the middle part: (a) D = 0.2; (b) D = 0.8.
Buildings 14 00070 g023
Figure 24. Damage to the hinge joint at the edge location: (a) overall; (b) detail.
Figure 24. Damage to the hinge joint at the edge location: (a) overall; (b) detail.
Buildings 14 00070 g024
Figure 25. Typical strain results for hinge joint damage developing from the girder’s edge: (a) D = 0.25; (b) D = 0.45.
Figure 25. Typical strain results for hinge joint damage developing from the girder’s edge: (a) D = 0.25; (b) D = 0.45.
Buildings 14 00070 g025
Figure 26. Cracks at the beam bottom: (a) longitudinal crack; (b) transverse crack.
Figure 26. Cracks at the beam bottom: (a) longitudinal crack; (b) transverse crack.
Buildings 14 00070 g026
Figure 27. Strain results under crack conditions: (a) longitudinal crack; (b) transverse crack.
Figure 27. Strain results under crack conditions: (a) longitudinal crack; (b) transverse crack.
Buildings 14 00070 g027
Figure 28. Bridge section and lane location.
Figure 28. Bridge section and lane location.
Buildings 14 00070 g028
Figure 29. Long-gauge sensor arrangement: (a) sensors in one group; (b) post location.
Figure 29. Long-gauge sensor arrangement: (a) sensors in one group; (b) post location.
Buildings 14 00070 g029
Figure 30. Sensor installation: (a) FBG sensors; (b) displacement meter.
Figure 30. Sensor installation: (a) FBG sensors; (b) displacement meter.
Buildings 14 00070 g030
Figure 31. Application of vehicle load: (a) location of static vehicle load; (b) static vehicle load; (c) moving vehicle load.
Figure 31. Application of vehicle load: (a) location of static vehicle load; (b) static vehicle load; (c) moving vehicle load.
Buildings 14 00070 g031
Figure 32. Typical strain results at 30 km/h: (a) South direction; (b) North direction.
Figure 32. Typical strain results at 30 km/h: (a) South direction; (b) North direction.
Buildings 14 00070 g032
Figure 33. Results of strain distribution: (a) South direction at 30 km/h; (b) North direction at 30 km/h; (c) South direction at 60 km/h; (d) North direction at 60 km/h.
Figure 33. Results of strain distribution: (a) South direction at 30 km/h; (b) North direction at 30 km/h; (c) South direction at 60 km/h; (d) North direction at 60 km/h.
Buildings 14 00070 g033
Table 1. Material properties of FE models.
Table 1. Material properties of FE models.
Elastic Mode for GirderCompressive Strength for GirderTensile Strength for GirderElastic Mode for JointsCompressive Strength for JointsTensile Strength for Girder
32.5 GPa19.1 Mpa1.71 Mpa30 Gpa14.3 Mpa1.43 Mpa
Table 2. Results of deflection of different girders.
Table 2. Results of deflection of different girders.
Girder No.Actual Deflection (mm)Deflection by Static Strain DistributionDeflection by Equivalent Strain Distribution
Deflection (mm)Error (%)Deflection (mm)Error (%)
10.830.863.390.885.63
20.830.852.610.863.34
30.850.850.000.850.62
40.850.850.000.860.84
50.860.860.000.860.11
60.860.860.000.86−0.05
Table 3. Results of deflection of different vehicle models.
Table 3. Results of deflection of different vehicle models.
Vehicle TypeActual
Deflection (mm)
Deflection by Static Strain
Distribution
Deflection by Equivalent Strain
Distribution
Deflection (mm)Error (%)Deflection (mm)Error (%)
Biaxial van0.860.860.000.86−0.03
Biaxial truck0.810.821.120.821.92
Three-axle truck1.441.461.351.450.81
Multi-axle trailer1.401.400.001.32−5.62
Table 4. Results of deflection at different vehicle speeds.
Table 4. Results of deflection at different vehicle speeds.
ItemActual
Deflection
Deflection by Static Strain
Distribution
Deflection by Equivalent Strain Distribution
30 km/h60 km/h90 km/h120 km/h
Deflection (mm)0.860.860.860.860.870.92
Error (%)/0.00−0.050.411.205.50
Table 5. Deflection results for hinge joint damage developing from the middle part.
Table 5. Deflection results for hinge joint damage developing from the middle part.
DActual Deflection (mm)Static Strain DistributionEquivalent Strain Distribution
Deflection (mm)Error (%)Deflection (mm)Error (%)
0.10.860.870.661.0015.88
0.20.890.890.011.0215.37
0.30.920.92−0.701.0917.63
0.40.970.95−1.451.1417.60
0.51.021.00−2.241.1815.97
0.61.081.04−3.111.2213.53
0.71.141.10−4.061.2610.46
0.81.221.16−5.141.306.91
0.91.311.23−6.341.352.78
1.01.761.65−6.331.66−5.78
Table 6. Deflection results for hinge joint damage developing from the girder’s edge.
Table 6. Deflection results for hinge joint damage developing from the girder’s edge.
DActual Deflection (mm)Static Strain DistributionEquivalent Strain Distribution
Deflection (mm)Error (%)Deflection (mm)Error (%)
0.050.870.881.120.881.54
0.100.870.881.100.881.51
0.150.870.881.070.881.42
0.200.870.881.040.881.27
0.250.870.881.000.880.94
0.300.870.880.960.870.36
0.350.870.880.950.86−0.60
0.400.870.881.000.85−2.18
0.450.880.891.130.84−4.05
Table 7. Deflection results under crack conditions.
Table 7. Deflection results under crack conditions.
ConditionActual Deflection (mm)Static Strain DistributionEquivalent Strain Distribution
Deflection (mm)Error (%)Deflection (mm)Error (%)
Longitudinal crack0.840.851.200.861.75
Transverse crack0.840.840.150.872.87
Table 8. Deflection results of on-site experiments.
Table 8. Deflection results of on-site experiments.
ConditionActual Deflection (mm)Static Strain DistributionEquivalent Strain Distribution
Deflection (mm)Error (%)Deflection (mm)Error (%)
South direction30 km/h0.710.68−3.400.756.02
60 km/h0.710.68−3.400.69−2.48
North direction30 km/h0.620.631.810.57−8.25
60 km/h0.620.631.810.59−4.25
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

Tang, Y.; Cang, J.; Zheng, B.; Tang, W. Deflection Monitoring Method for Simply Supported Girder Bridges Using Strain Response under Traffic Loads. Buildings 2024, 14, 70. https://doi.org/10.3390/buildings14010070

AMA Style

Tang Y, Cang J, Zheng B, Tang W. Deflection Monitoring Method for Simply Supported Girder Bridges Using Strain Response under Traffic Loads. Buildings. 2024; 14(1):70. https://doi.org/10.3390/buildings14010070

Chicago/Turabian Style

Tang, Yongsheng, Jigang Cang, Bohan Zheng, and Wei Tang. 2024. "Deflection Monitoring Method for Simply Supported Girder Bridges Using Strain Response under Traffic Loads" Buildings 14, no. 1: 70. https://doi.org/10.3390/buildings14010070

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