Next Article in Journal
Reducing Dynamic Power Consumption in Mixed-Critical Real-Time Systems
Next Article in Special Issue
Rail Diagnostics Based on Ultrasonic Guided Waves: An Overview
Previous Article in Journal
In Silico-Enhanced Treatment and Rehabilitation Planning for Patients with Musculoskeletal Disorders: Can Musculoskeletal Modelling and Dynamic Simulations Really Impact Current Clinical Practice?
Previous Article in Special Issue
Mooring-Failure Monitoring of Submerged Floating Tunnel Using Deep Neural Network
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

State-of-the-Art Review on Determining Prestress Losses in Prestressed Concrete Girders

1
Bridge Engineering Division, National Center for Research on Earthquake Engineering (NCREE), Taipei 10668, Taiwan
2
Department of Civil Engineering, National Taiwan University, Taipei 10617, Taiwan
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(20), 7257; https://doi.org/10.3390/app10207257
Submission received: 25 August 2020 / Revised: 29 September 2020 / Accepted: 14 October 2020 / Published: 16 October 2020
(This article belongs to the Special Issue Novel Approaches for Structural Health Monitoring)

Abstract

:
Prestressing methods were used to realize long-span bridges in the last few decades. For their predictive maintenance, devices and dynamic nondestructive procedures for identifying prestress losses were mainly developed since serviceability and safety of Prestressed Concrete (PC) girders depend on the effective state of prestressing. In fact, substantial long term prestress losses can induce excessive deflections and cracking in large span PC bridge girders. However, old unsolved problematics as well as new challenges exist since a variation in prestress force does not significantly affect the vibration responses of such PC girders. As a result, this makes uncertain the use of natural frequencies as appropriate parameters for prestress loss determinations. Thus, amongst emerging techniques, static identification based on vertical deflections has preliminary proved to be a reliable method with the goal to become a dominant approach in the near future. In fact, measured vertical deflections take accurately and instantaneously into account the changes of structural geometry of PC girders due to prestressing losses on the equilibrium conditions, in turn caused by the combined effects of tendon relaxation, concrete creep and shrinkage, and parameters of real environment as, e.g., temperature and relative humidity. Given the current state of quantitative and principled methodologies, this paper represents a state-of-the-art review of some important research works on determining prestress losses conducted worldwide. The attention is principally focused on a static nondestructive method, and a comparison with dynamic ones is elaborated. Comments and recommendations are made at proper places, while concluding remarks including future studies and field developments are mentioned at the end of the paper.

1. Introduction

The first applications of prestressing methods to concrete structures go back to the first half of the 20th century. Today prestressing is widely used for many applications, ranging from small members, such as railway sleepers, to more important structures such as bridges, long and light precast flooring, and roofing elements for constructions. Serviceability and safety of Prestressed Concrete (PC) structures rely on the effective state of prestress force [1]. In fact, prestressing methods are principally utilized to reduce deflections and to partially counterbalance the effect of dead and live loads in the case of bridges [2]. As a result, an extreme loss of prestressing may cause excessive deflections or jeopardize the performance of large span PC girders by indicating cracking phenomena [3]. For this reason, devices and dynamic approaches capable of determining prestress losses largely developed. In particular, dynamic Structural Health Monitoring (SHM) techniques also generated for damage identification purposes based on the vibration responses of the span girders, so preventing maintenance, repair, or replacement of a bridge [4,5,6]. Damage detection techniques as well developed using different equipment and methods as, e.g., parked vehicles inducing frequency variation [7], long-gauge Fiber Bragg Grating (FBG) [8], or hybrid vibration testing data [9]. Therefore, the operating state of bridges can be controlled. However, prestress losses can be directly, simply, and accurately determined over time if the internal tendons of PC girders are instrumented by load cells, vibrating wire strain gauges, or elasto-magnetic sensors during construction [10,11,12,13]. Besides, FBG sensors can be embedded in seven-wire strands along PC girders for long term monitoring of tensile forces [14,15]. Although instrumentation of external tendons is easy during their serviceability, NonDestructive Testing (NDT) methods are required. Nevertheless, as far as the influence of prestressing on the dynamics of PC girders is concerned, the discussion is still ongoing.
Several procedures and equations are available in design codes for predicting prestress losses. According to ACI 318-2019 [16] and PCI DH [17], reasonable estimations can be calculated. For unusual design conditions and special structures, a more detailed procedure established by PCI CPL [18] can be considered. AASHTO LRFD [19] adopted new procedures since the previous prestress loss methods led to unrealistic applications with high strength concrete. However, AASHTO Standard specifications [20] remain in accordance with AASHTO LRFD [21]. PCI BDM [22] includes both AASHTO Standard [20] and LRFD [21] methods.
A series of studies were conducted to measure prestress losses in PC girders, and to compare them versus design code estimations. Among these works, there are laboratory tests of old PC girders removed from existing bridges, and experiments including fabrication, testing, and field monitoring of PC members under service. Table 1 in [23] and Table 6 in [24] summarize an extensive literature review on references, PC member identification (type, old time), testing place, experimental technique used, time of study, and measured losses. As observed in these tables, measured prestress losses exceeded those predicted by design codes in some cases. On the other hand, measured prestress losses which were in line with the values expected by codes were obtained in PC girders, which exceeded the allowable compressive stress limit [25]. To overcome this problem, Caro et al. [26] used the ECADA+ method [27] to measure the effective prestress forces in a number of PC specimens for over 1 year and, consequently, compared the results with prestress losses estimated by several codes. Although design code-based predictions can be considered as quite satisfactory, they are very conservative [28]. Accordingly, there are difficulties in determining prestress losses related to factors including, inter alia, assumptions about the properties of prestressing systems and time-dependent phenomena, such as long term degradation processes, tendon relaxation, creep and shrinkage of concrete, and parameters of the real environment [29,30].
Given the current state of quantitative and principled methodologies, this paper represents a state-of-the-art review of some important research works conducted worldwide on determining prestress losses in PC girders. At first, laboratory, numerical investigations, and testing methods are reviewed. Secondly, the article focuses on a static NDT method, and a comparison with dynamic ones is elaborated, since old unsolved problematics as well as new challenges exist because a variation in prestress force does not significantly influence the vibration responses of PC girders. Consequently, this makes uncertain the use of natural frequencies as appropriate parameters for prestress loss determinations. Thus, amongst emerging techniques, static identification based on vertical deflections has preliminary proved to be a reliable method with the aim to become a dominant testing approach in the near future. In fact, measured vertical deflections take accurately and instantaneously into account the changes of structural geometry of PC girders due to prestressing losses on the equilibrium conditions, in turn caused by the combined effects of relaxation of tendon, concrete creep and shrinkage, and parameters of the real environment as, e.g., temperature and relative humidity. Comments and recommendations are made at proper places, whilst concluding remarks including future investigations and field developments are mentioned at the end of the article.

2. Works Conducted by Researchers Worldwide

Research was devoted to analyzing the influence of prestress force on the dynamic behavior of PC girders. A state-of-the-art review of the primary contributions on this topic is presented herein, including laboratory, numerical investigations, and testing methods.

2.1. Laboratory Investigations

Hop [31] monitored the vibration responses of a number of PC beams. Investigation focused on the influence of prestress forces on frequency and damping. The author found that applying an increase in prestress force, acting unevenly on the member, would increase the vibration frequency. In many cases, it was measured that the application of further levels of prestressing increase would result in drop of vibration frequency.
Similar experimental results were obtained by Saiidi et al. [32] on a PC girder with concentric tendon. The research proved an increase of the first eigenfrequency from 11.41 Hz for the case of null prestressing, to 15.07 Hz for the maximum magnitude of prestressing. Besides, the authors observed that an increase in prestressing seems to influence microcrack closure and, consequently, increment stiffness and natural frequencies of PC beams.
Miyamoto et al. [33] tested the dynamic response of PC girders, strengthened with external tendons. According to their results, prestress forces applied to external tendons influence the frequency vibrations of PC girders.
Lu and Law [34] tested a PC beam with a straight concentric tendon. Two conditions were studied, id est, with and without the prestress force of 66.7 kN. The authors noted that the prestressing induced an increment in the first three eigenfrequencies within a range of 0.4–2.1%.
Xiong and Zhang [35] tested three simply supported PC girders with different configurations of external tendons. The authors observed that the natural frequency initially increased with the increase in prestress force. Vice versa, the natural frequency decreased after cracks induced by prestressing.
Kim et al. [36] tested a PC girder with many damage scenarios of prestress losses. Starting from a state of absence of prestress losses, the prestress force was then gradually reduced to zero. During this unloading process, vibration measurements allowed to estimate reductions of the first four eigenfrequencies up to values of 4.0–4.4% from the initial stage to the final one.
Jang et al. [37] tested a number of six PC beams with a bonded tendon. By applying continuously an increase of prestressing from 0 to 523 kN, the authors observed a progressive increase of the first eigenfrequency from 7.6 to 8.7 Hz.
Noh et al. [38] performed tests on three PC girders with different configurations of tendons. The researchers detected that the natural frequency generally increases as tension force in prestressing steel increases. Additionally, they observed that natural frequencies of PC girders are affected by other parameters, such as beam camber, geometric stiffness of tendon, and stiffness effect of beam-tendon system.
The results of the aforementioned relevant works, id est, of Hop [31], Saiidi et al. [32], Kim et al. [36], and Jang et al. [37], declared that for lower values of prestress force, an increase in prestressing generates an increase in eigenfrequencies, especially for the fundamental frequency. Conversely, for higher levels of prestressing, the rate of increase of the eigenfrequencies tends to decrease. Moreover, changes in vibration frequencies were observed to be higher for smaller eccentricities of prestressing tendon. By considering such investigations, it can generally be claimed that prestress force slightly affects the dynamics of PC girders (Table 1). Nevertheless, relevance of these modifications depends on many characteristics (e.g., cracking and nonlinearity of concrete, bonding and eccentricity of tendon) which generate counterbalancing effects making it difficult to determine an evident relation between dynamic properties and prestress force.

2.2. Numerical Investigations

Numerical studies were conducted by addressing the effect of prestress force on the dynamics of PC girders. During simulation of a moving force identification method which assumed the effects of prestressing, Chan and Yung [39] found that the natural frequencies of a PC bridge decrease with an increase in prestress force. This is well-known as the “compression-softening” effect and, in general, occurs in Euler–Bernoulli beams and PC members preserved against crack formation [40,41,42].
Kim et al. [43] investigated prestress loss identifications in PC girders based on the measurement variations of natural frequencies. Comparison between the experimental results obtained by Saiidi et al. [32] and the previsions of their model validated their method.
Law and Lu [44] analyzed the time-domain response of a PC girder under dynamic excitation. By comparing their results of numerical simulations with theoretical findings, the authors identified the prestress force in the time domain by displacement and strain measurements. According to their findings, the natural frequencies decrease as the prestressing increases.
Hamed and Frostig [45] developed a nonlinear solution for modeling the behavior of PC girders with a bonded or an unbonded tendon. Based on the derived governing equations, the authors proved that prestressing does not influence the natural frequencies of PC girders.
Jaiswal [46] underlined that the increase of a PC girder’s stiffness (and frequency) depends on the tendon’s eccentricity, thus inducing greater moment and stiffening effect along the member.
Limongelli et al. [47] investigated the prediction of early warning signs of deterioration in a PC beam caused by prestress losses. The investigators pointed out that vibration frequencies of PC girders significantly change only under the effects of crack initiation or crack re-opening.
Gan et al. [48] validated the experiments done by Jang et al. [37] and Noble et al. [49] by a Finite Element (FE) model, where the influence of prestressing on the natural frequencies was simulated by the existence of early-age shrinkage cracks inside the concrete.
Bonopera et al. [50,51] found that the fundamental frequency of uncracked PC girders with a parabolic tendon is unaffected by the prestress force because the course of the “compression–softening” theory being cancelled out by the increase of elastic modulus due to the concrete’s consolidation/hardening with time [52].
Luna Vera et al. [53] investigated the flexural performance of two PC girders under uncracked and cracked conditions for further applications of SHM. In relation to the uncracked state, the authors affirmed that changes in natural frequency due to prestress losses are negligible.
Looking at the relevant works above mentioned, id est, at Hamed and Frostig [45], Jaiswal [46], Limongelli et al. [47], and Luna Vera et al. [53], no significant agreement between the effect of prestress force and dynamics of uncracked PC girders was observed (Table 2). As a result, natural frequencies are generally considered inappropriate parameters for determining prestress losses, as also indicated by Saiidi et al. [32] and Bonopera et al. [50,51].

2.3. Testing Methods

Testing methods are in general required for estimating prestress losses, and include five typologies (Table 3): (1) static load testing to determine crack initiation or crack re-opening loads to obtain the available compressive stress in the bottom flange of a PC girder [23,28,54,55]. (2) Severing the prestressing tendon by cutting it into a representative exposed length after placing strain gauges on the tendon [23,56]. (3) Relating the tension in the tendon to a vertical deflection recorded when known weights are suspended from it on a representative exposed length [54,57]. (4) Determining the side pressure to close the induced crack in a small cylindrical hole drilled adjacent to the tendon in the bottom flange of a PC girder [55,58]. (5) Vibration testing to determine the natural frequencies and/or dynamic responses of a PC girder [59,60,61,62,63,64,65].
Methods 1 and 2 are destructive approaches which cause damages. In Method 1, a large amount of vertical loads must be applied to involve crack initiation, or to force a main crack to form in the same location as a monitored crack.
Method 3 is a semidestructive test for exposed tendons, and involves accurately determining the exposed length for calculations.
Method 4 is a NDT approach for embedded tendons which involves a negligible local damage along the PC span girder. This is not possible when the tendon is too close to the concrete surface because it could be damaged by the deep hole-cut. Notably, further studies in a controlled environment with monitoring of actual prestress forces were suggested in order to validate its applicability for PC girders with a parabolic tendon [55].
Method 5 represents a series of NDT techniques which require local vibration measurements along the PC girder. With regards, Law et al. [66], Li et al. [67], and Xiang et al. [68] performed numerical simulations using the dynamics of PC girders to moving vehicular loads. In their methods, prestress force identifications were obtained using vibration measures if the prestress force in the tendon is assumed as equivalent to an external compressive force applied to the beam ends [69,70,71,72,73,74,75,76,77,78,79,80,81,82]. Consequently, the natural frequency of the PC girder tends to increase with a decrease in prestress force according to the “compression–softening” theory. Notably, whether prestressed beams are subjected to the “compression–softening” effect was extensively discussed as, e.g., in the literature review by Noble et al. [49,83]. Most of the dynamic methods, cited therein, are based on the identified modal characteristics of a PC girder, id est, natural frequencies and/or mode shapes. The modal characteristics depend on the PC girder’s stiffness and, accordingly, are affected by the prestress force. Particularly, such vibration-based identifications require an accurate selection of the mode shape to be utilized in the procedures. In fact, selecting the optimal frequency a priori is challenging, and different frequencies yield varying degrees of accuracy in prestress force determinations.

3. Static NDT Methods

Amongst vibration-based techniques, static identifications using vertical deflections have proved to be reliable methods for axial force identification in beam elements. Indeed, measured vertical deflections take accurately into account the changes of structural geometry of the members due to axial force variations on the equilibrium conditions [40,41,71,74,77,84] (Figure 1a,b). Experimental simulations were as well conducted on members belonging to space frames and trusses [78,79,85]. Likewise, Bonopera et al. [80] verified the feasibility of estimating prestress force in a PC girder using vertical deflections measured by three-point bending tests. It is also worth noting that this approach only adopts static parameters, thus, in contrast to vibration-based techniques, does not require selecting experimental data for use in the algorithms.

3.1. Brief on Works Conducted by Bonopera et al. (2018)

The static approach was originally developed for detecting axial force in compressed steel beams [77]. Subsequently, it was employed for identifying prestress forces in PC girders [80]. In this last case, the reference model comprised a simply supported Euler–Bernoulli beam of 250 mm in width, 400 mm in height, and length L of 6.62 m, made with a high strength concrete, and prestressed by a straight unbonded tendon, where the prestress force N was assumed as an external compressive force eccentrically applied to the end constraints N e (Figure 2a). The cross sectional second moment of the area of the PC beam’s section I was equal to 1.3333 × 109 mm4. Besides, a bending deflection v(1) along the aforementioned beam’s model, of 0.01 mm in accuracy, was properly approximated by multiplying the corresponding first-order deflection by the “magnification factor” of the second-order effects, id est, according to the “compression-softening” theory [40,41,42] (Figure 2b).
Experiments on a PC beam specimen, having the test configuration above mentioned (Figure 2a,b), were arranged in the laboratory of the National Center for Research on Earthquake Engineering (NCREE) of Taipei, Taiwan [80], where a research program, based on testing uncracked PC bridge member prototypes, began in 2015 (Figure 3). All geometrical dimensions were checked by laser rangefinder and caliper, of 0.01 mm in tolerance, after the PC beam was fixed on the simple supports. First, deflected-shape measurements v(1) along its span, obtained from 27 three-point bending tests with different applied prestress forces N, measured by a load cell placed at both end constraints [86], were examined to verify the accuracy of the assumptions of the beam’s model (Figure 2b). The span of the PC beam specimen was specifically instrumented to short term measure such static deflection shapes, with an accuracy of 0.01 mm, by a set of displacement transducers. Temperature and relative humidity of real environment of the PC beam were not continuously recorded during testing. Second, based on the “magnification factor” formula of the “compression-softening” theory [40,41,42], prestress force determinations were obtained using two series of vertical deflections v(1), id est, those recorded at the quarter v2(1) and at the midspan v4(1) of the PC beam, respectively. Information regarding the flexural rigidity of the PC beam were in addition required. In detail, average values of the chord elastic modulus Eaver of the high strength concrete, with an accuracy of 1 MPa, were estimated by compression tests on a series of cylinders cast at the same time of the PC beam [87] to determine its increment caused by the concrete’s consolidation/hardening with time [52].

3.2. Prestress Force Determinations Obtained by Bonopera et al. (2018)

Table 4 lists the prestress force determinations Na obtained using the vertical deflections v2(1) and corresponding experimental values Ψ = F L3/Eaver I in Equation (8a) (Test 1), as well as the vertical deflections v4(1) and corresponding parameters Ψ in Equation (8b) (Test 2), respectively. Equations (8a) and (8b) are illustrated in Bonopera et al. [80]. In specific, the nine test combinations represent the best prestress force determinations among the three test repetitions performed (Section 3.1). The chord elastic modulus Eaver of the high strength concrete, for each day of execution of the experiments (Section 3.1), was utilized as parameter in the identification process. The corresponding first-order vertical deflections were instead estimated by Equations (4a) and (4b), as similarly reported in Bonopera et al. [80]. Table 4 additionally shows the related percentage errors Δ = (NaN) / N. In general, poor estimates Na were obtained when prestressing N equal to 617 and 620 kN were assigned (Figure 2a,b). Vice versa, the test combinations with prestressing that induced second-order effects greater than 6.5%, id est, when N ≧ 721 kN, furnished excellent identifications of prestress forces Na. In fact, in this last case, estimation errors were lower, in absolute value, than 6.1%.
Sensitivity analyses were elaborated for the prestress force determinations based on Equations (8a) and (8b). The vertical deflections v2(1) and v4(1), calculated with 0.01 mm in accuracy by Equations (3a) and (3b) reported in Bonopera et al. [80], and parameter Ψ were modified to generate possible experimental errors. In detail, deflections v2(1), v4(1), and Ψ were alternatively multiplied by 0.99 and 1.01 to reproduce 14 combinations of simulated values for nine different assumed prestress forces N. The average value of the applied vertical loads, Faver = 22.6 kN (Figure 2b), was taken into account in the manipulations. Figure 4 depicts a comparison between the worst determined Na and assumed values N conducted using vertical deflections v2(1) and v4(1), both of which yielded a constant error of about ±107 kN. Based on all the results obtained, a favorable correspondence between analytical N and experimental determinations of prestress force Na was found when midspan deflections v4(1) were taken into account.

4. Concluding Remarks

A state-of-the-art review of some important research works conducted worldwide on determining prestress losses in PC girders allowed to analyze various information and trace future developments. Some references affirm that a variation in prestress force does not significantly affect the vibration responses of PC girders. Accordingly, this makes uncertain the use of natural frequencies as appropriate parameters for prestress loss determinations. Vice versa, most of laboratory works show a slight increment in the eigenfrequencies under the increase in prestress force. This behavior is related to the concrete mechanics, and is a main consequence of the effect of crack and microcrack closure along PC girders. However, vibration-based identification methods require an accurate selection of the mode shape because different natural frequencies provide varying degrees of accuracy in prestress force evaluations.
By considering these characteristics, the manuscript focused then on the static NDT method preliminary proposed by Bonopera et al. [80] through laboratory simulations on an uncracked PC beam specimen. The procedure can accurately and instantaneously determine the effective prestress force using vertical deflection measurements, of 0.01 mm in accuracy, under actual ambient conditions [51,81]. The precision of estimation improved when the PC beam was subjected to a high prestress force and, moreover, when midspan deflections were taken into account. Information regarding the flexural rigidity of a PC girder under investigation are also necessary. With regards, an average value of the chord elastic modulus, of 1 MPa in accuracy, must be estimated by compression tests on a set of concrete cores, drilled along its span, at the time of deflection measurements. Besides, the static NDT method does not require any direct measure of the tension force in the tendon and, mostly, in contrast to dynamic NDT ones, does not request selecting experimental data for use in the algorithms (Table 5).
In conclusion, to make the NDT method applicable in situ, further studies should focus on the measurement of vertical deflections induced by bending tests with vehicle loading along PC bridge girders [88,89,90], in which their constraint stiffness should be evaluated with unknown boundary conditions. In fact, static vertical deflections take accurately and instantaneously into account the changes of structural geometry due to prestressing losses on the equilibrium conditions [51,80,81], in turn caused by the combined effects of tendon relaxation, concrete creep and shrinkage, and parameters of real environment of the PC girder as, e.g., temperature and relative humidity [49] (Figure 3a). The FBG-DSM liquid-level system [77,91] is instead an effective measurement device because it can provide bridge deflections over long distances up to 0.01 mm in accuracy (Figure 3b), referred to an absolute point, without any external physical reference and requirements of good environmental conditions, accessibility and visibility in situ.

Author Contributions

Methodology, M.B.; software, Z.-K.L.; validation, M.B., K.-C.C. and Z.-K.L.; formal analysis, M.B.; investigation, M.B. and K.-C.C.; resources, Z.-K.L.; data curation, M.B. and Z.-K.L.; writing—original draft preparation, M.B.; writing—review and editing, M.B.; visualization, M.B. and Z.-K.L.; supervision, K.-C.C.; project administration, M.B.; funding acquisition, M.B. and K.-C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Applied Research Laboratories Project of Taiwan (grant number: NCREE–06109A1700), and Ministry of Science and Technology (MOST) of Taiwan in the framework of the project “Recruitment of Visiting Science and Technology Personnel with MOST Funding” (grant number: MOST 108–2811–E–492–001).

Acknowledgments

The authors would like to thank the National Applied Research Laboratories Project of Taiwan (NCREE–06109A1700), and Ministry of Science and Technology (MOST) of Taiwan in the framework of the project “Recruitment of Visiting Science and Technology Personnel with MOST Funding” (MOST 108–2811–E–492–001) for their financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Breccolotti, M.; Materazzi, A. Prestress Losses and Camber Growth in Wing-Shaped Structural Members. PCI J. 2015, 60, 98–117. [Google Scholar] [CrossRef]
  2. Robertson, I.N. Prediction of vertical deflections for a long-span prestressed concrete bridge structure. Eng. Struct. 2005, 27, 1820–1827. [Google Scholar] [CrossRef]
  3. Huang, H.; Huang, S.-S.; Pilakoutas, K. Modeling for assessment of long-term behavior of prestressed concrete box-girder bridges. J. Bridg. Eng. 2018, 23, 04018002. [Google Scholar] [CrossRef] [Green Version]
  4. Domaneschi, M.; Limongelli, M.; Martinelli, L. Multi-site damage localization in a suspension bridge via aftershock monitoring. Ing. Sismica 2013, 30, 56–72. [Google Scholar]
  5. Domaneschi, M.; Limongelli, M.; Martinelli, L. Damage detection and localization on a benchmark cable-stayed bridge. Earthq. Struct. 2015, 8, 1113–1126. [Google Scholar] [CrossRef]
  6. Limongelli, M.P.; Tirone, M.; Surace, C. Non-destructive monitoring of a prestressed bridge with a data-driven method. Proc. SPIE Int. Soc. Opt. Eng. 2017, 10170, 1017033. [Google Scholar]
  7. He, W.-Y.; Ren, W.-X. Structural damage detection using a parked vehicle induced frequency variation. Eng. Struct. 2018, 170, 34–41. [Google Scholar] [CrossRef]
  8. Chen, S.Z.; Wu, G.; Feng, D.C. Damage detection of highway bridges based on long-gauge strain response under stochastic traffic flow. Mech. Syst. Signal Process. 2019, 127, 551–572. [Google Scholar] [CrossRef]
  9. Chen, G.W.; Chen, X.; Omenzetter, P. Modal parameter identification of a multiple-span post-tensioned concrete bridge using hybrid vibration testing data. Eng. Struct. 2020, 219, 110953. [Google Scholar] [CrossRef]
  10. Saiidi, M.; Shields, J.; O’Connor, D.; Hutchens, E. Variation of prestress forces in a prestressed concrete bridge during the first 30 months. PCI J. 1996, 41, 66–72. [Google Scholar] [CrossRef]
  11. Saiidi, M.; Hutchens, E.; Gardella, D. Bridge prestress losses in dry climate. J. Bridge Eng. 1998, 3, 111–116. [Google Scholar] [CrossRef]
  12. Guo, T.; Chen, Z.; Lu, S.; Yao, R. Monitoring and analysis of long-term prestress losses in post-tensioned concrete beams. Measurement 2018, 122, 573–581. [Google Scholar] [CrossRef]
  13. Kim, J.; Kim, J.W.; Park, S. Investigation of applicability of an embedded EM sensor to measure the tension of a PSC girder. J. Sens. 2019, 2019, 2469647. [Google Scholar] [CrossRef]
  14. Cho, K.; Kim, S.T.; Cho, J.R.; Park, Y.H. Estimation of tendon force distribution in prestressed concrete girders using smart strand. Appl. Sci. 2017, 7, 1319. [Google Scholar] [CrossRef] [Green Version]
  15. Sung, H.J.; Do, T.M.; Kim, J.M.; Kim, Y.S. Long-term monitoring of ground anchor tensile forces by FBG sensors embedded tendon. Smart. Struct. Syst. 2017, 19, 269–277. [Google Scholar] [CrossRef]
  16. ACI (American Concrete Institute). Building Code Requirements for Structural Concrete and Commentary; ACI: Farmington Hills, MI, USA, 2019; pp. 318–319. [Google Scholar]
  17. PCI. Design Handbook, 7th ed.; Precast/Prestressed Concrete Institute: Chicago, IL, USA, 2010. [Google Scholar]
  18. PCI Commitee on Prestress Losses. Recommendations for estimating prestress losses. PCI J. 1975, 20, 43–75. [Google Scholar]
  19. AASHTO (American Association of State Highway and Transportation Officials). AASHTO LRFD Bridge Design Specifications, 8th ed.; American Association of State Highway and Transportation Officials: Washington, DC, USA, 2017. [Google Scholar]
  20. AASHTO (American Association of State Highway and Transportation Officials). AASHTO Standard Specifications for Highways Bridges, 17th ed.; American Association of State Highway and Transportation Officials: Washington, DC, USA, 2002. [Google Scholar]
  21. AASHTO (American Association of State Highway and Transportation Officials). AASHTO LRFD Bridge Design Specifications, 3rd ed.; American Association of State Highway and Transportation Officials: Washington, DC, USA, 2004. [Google Scholar]
  22. PCI. Bridge Design Manual, 3rd ed.; Precast/Prestressed Concrete Institute: Chicago, IL, USA, 2011. [Google Scholar]
  23. Baran, E.; Shield, C.K.; French, C.E. A comparison of methods for experimentally determining prestress losses in pretensioned prestressed concrete girders. In Historic Innovations in Prestressed Concrete; Russell, B.W., Gross, S.P., Eds.; American Concrete Institute: Farmington Hills, MI, USA, 2005; pp. 161–179. [Google Scholar]
  24. Al-Omaishi, N.; Tadros, M.K.; Seguirant, S.J. Estimating prestress loss in pretensioned, high-strength concrete members. PCI J. 2009, 54, 132–159. [Google Scholar] [CrossRef]
  25. Hale, W.M.; Russell, B.W. Effect of allowable compressive stress at release on prestress losses and on the performance of precast, prestressed concrete bridge girders. PCI J. 2006, 51, 14–25. [Google Scholar] [CrossRef]
  26. Caro, L.A.; Martí-Vargas, J.R.; Serna, P. Prestress losses evaluation in prestressed concrete prismatic specimens. Eng. Struct. 2013, 48, 704–715. [Google Scholar] [CrossRef]
  27. Martí-Vargas, J.R.; Caro, L.A.; Serna, P. Experimental technique for measuring the long-term transfer length in prestressed concrete. Strain 2013, 49, 125–134. [Google Scholar] [CrossRef]
  28. Garber, D.B.; Gallardo, J.M.; Deschenes, D.J.; Bayrak, O. Experimental investigation of prestress losses in full-scale bridge girders. ACI Struct. J. 2015, 112, 553–564. [Google Scholar] [CrossRef] [Green Version]
  29. Bažant, Z.P.; Jirásek, M.; Hubler, M.H.; Carol, I. RILEM draft recommendation: TC-242-MDC multi-decade creep and shrinkage of concrete: Material model and structural analysis. Model B4 for creep, drying shrinkage and autogenous shrinkage of normal and high-strength concretes with multi-decade applicability. Mater. Struct. 2015, 48, 753–770. [Google Scholar]
  30. Kolínský, V.; Vítek, J.L. New model for concrete creep and shrinkage prediction and its application. Solid State Phenom. 2016, 249, 125–130. [Google Scholar] [CrossRef]
  31. Hop, T. The effect of degree of prestressing and age of concrete beams on frequency and damping of their free vibration. Mater. Struct. 1991, 24, 210–220. [Google Scholar] [CrossRef]
  32. Saiidi, M.; Douglas, B.; Feng, S. Prestress force effect on vibration frequency of concrete bridges. J. Struct. Eng. 1994, 120, 2233–2241. [Google Scholar] [CrossRef]
  33. Miyamoto, A.; Tei, K.; Nakamura, H.; Bull, J.W. Behavior of prestressed beam strengthened with external tendons. J. Struct. Eng. 2000, 126, 1033–1044. [Google Scholar] [CrossRef]
  34. Lu, Z.; Law, S. Identification of prestress force from measured structural responses. Mech. Syst. Signal Pr. 2006, 20, 2186–2199. [Google Scholar] [CrossRef]
  35. Xiong, H.X.; Zhang, Y.T. Theoretical analysis of natural frequency of externally prestressed concrete beam based on rigidity correction. Acad. J. Xian Jiaotong Univ. 2009, 21, 31–35. [Google Scholar]
  36. Kim, J.T.; Park, J.H.; Hong, D.S.; Park, W.S. Hybrid health monitoring of prestressed concrete girder bridges by sequential vibration-impedance approaches. Eng. Struct. 2010, 32, 115–128. [Google Scholar] [CrossRef]
  37. Jang, J.; Lee, H.; Hwang, K.; Song, Y. A sensitivity analysis of the key parameters for the prediction of the pre-stress force on bonded tendons. Nucl. Eng. Technol. 2010, 42, 319–328. [Google Scholar] [CrossRef] [Green Version]
  38. Noh, M.H.; Seong, T.R.; Lee, J.; Park, K.S. Experimental investigation of dynamic behavior of prestressed girders with internal tendons. Int. J. Steel Struct. 2015, 15, 401–414. [Google Scholar] [CrossRef]
  39. Chan, T.H.T.; Yung, T.H. A theoretical study of force identification using prestressed concrete bridges. Eng. Struct. 2000, 22, 1529–1537. [Google Scholar] [CrossRef]
  40. Timoshenko, S.P.; Gere, J.M. Theory of Elastic Stability; McGraw-Hill: New York, NY, USA, 1961. [Google Scholar]
  41. Bažant, Z.P.; Cedolin, L. Stability of Structures; Oxford University Press: New York, NY, USA, 1991. [Google Scholar]
  42. Young, W.C.; Budynas, R.G. Dynamic and Temperature Stresses. In Roark’s Formulas for Stress and Strain; Young, W.C., Budynas, R.G., Eds.; McGraw–Hill: New York, NY, USA, 2002; pp. 767–768. [Google Scholar]
  43. Kim, J.T.; Yun, C.B.; Ryu, Y.S.; Cho, H.M. Identification of prestress-loss in PSC beams using modal information. Struct. Eng. Mech. 2004, 17, 467–482. [Google Scholar] [CrossRef]
  44. Law, S.; Lu, Z. Time domain responses of a prestressed beam and prestress identification. J. Sound Vib. 2005, 288, 1011–1025. [Google Scholar] [CrossRef]
  45. Hamed, E.; Frostig, Y. Natural frequencies of bonded and unbonded prestressed beams-prestress force effects. J. Sound Vib. 2006, 295, 28–39. [Google Scholar] [CrossRef]
  46. Jaiswal, O.R. Effect of prestressing on the first flexural natural frequency of beams. Struct. Eng. Mech. 2008, 28, 515–524. [Google Scholar] [CrossRef]
  47. Limongelli, M.P.; Siegert, D.; Merliot, E.; Waeytens, J.; Bourquin, F.; Vidal, R.; Le Corvec, V.; Gueguen, I.; Cottineau, L.M. Damage detection in a post tensioned concrete beam—Experimental investigation. Eng. Struct. 2016, 128, 15–25. [Google Scholar] [CrossRef] [Green Version]
  48. Gan, B.Z.; Chiew, S.P.; Lu, Y.; Fung, T.C. The effect of prestressing force on natural frequencies of concrete beams—A numerical validation of existing experiments by modelling shrinkage crack closure. J. Sound Vib. 2019, 455, 20–31. [Google Scholar] [CrossRef]
  49. Noble, D.; Nogal, M.; O’Connor, A.; Pakrashi, V. The effect of prestress force magnitude and eccentricity on the natural bending frequencies of uncracked prestressed concrete beams. J. Sound Vib. 2016, 365, 22–44. [Google Scholar] [CrossRef] [Green Version]
  50. Bonopera, M.; Chang, K.C.; Chen, C.C.; Sung, Y.C.; Tullini, N. Experimental study on the fundamental frequency of prestressed concrete bridge beams with parabolic unbonded tendons. J. Sound Vib. 2019, 455, 150–160. [Google Scholar] [CrossRef]
  51. Bonopera, M.; Chang, K.C.; Lin, T.K.; Tullini, N. Influence of prestressing on the behavior of uncracked concrete beams with a parabolic bonded tendon. Struct. Eng. Mech. 2021, 77. Accepted. [Google Scholar]
  52. Singh, B.P.; Yazdani, N.; Ramirez, G. Effect of a time dependent concrete modulus of elasticity on prestress losses in bridge girders. Int. J. Concr. Struct. Mater. 2013, 7, 183–191. [Google Scholar] [CrossRef] [Green Version]
  53. Luna Vera, O.S.; Oshima, Y.; Kim, C.W. Flexural performance correlation with natural bending frequency of post-tensioned concrete beam: Experimental investigation. J. Civ. Struct. Health Monit. 2020, 10, 135–151. [Google Scholar] [CrossRef]
  54. Labia, Y.; Saiidi, M.S.; Douglas, B. Full-scale testing and analysis of 20-year-old pretensioned concrete box girders. ACI Struct. J. 1997, 94, 471–482. [Google Scholar]
  55. Bagge, N.; Nilimaa, J.; Elfgren, L. In-situ methods to determine residual prestress forces in concrete bridges. Eng. Struct. 2017, 135, 41–52. [Google Scholar] [CrossRef]
  56. Halsey, J.T.; Miller, R. Destructive testing of two forty-year-old prestressed concrete bridge beams. PCI J. 1996, 41, 84–93. [Google Scholar] [CrossRef]
  57. Civjan, S.A.; Jirsa, J.O.; Carrasquillo, R.L.; Fowler, D.W. Instrument to evaluate remaining prestress in damaged prestressed concrete bridge girders. PCI J. 1998, 43, 62–69. [Google Scholar] [CrossRef]
  58. Azizinamini, A.; Keeler, B.J.; Rohde, J.; Mehrabi, A.B. Application of a new nondestructive evaluation technique to a 25-year-old prestressed concrete girder. PCI J. 1996, 41, 82–95. [Google Scholar] [CrossRef]
  59. Kim, J.T.; Ryu, Y.S.; Yun, C.B. Vibration based method to detect pre–stress loss in beam type bridges. Smart Struct. Mater. 2003, 5057, 559–569. [Google Scholar] [CrossRef]
  60. Bu, J.Q.; Wang, H.Y. Effective pre–stress identification for a simply supported PRC beam bridge by BP neural network method. J. Vib. Shock 2011, 30, 155–159. [Google Scholar]
  61. Xu, J.; Sun, Z. Pre–stress force identification for eccentrically prestressed concrete beam from beam vibration response. Technol. Sci. Press SL 2011, 5, 107–115. [Google Scholar]
  62. Shi, L.; He, H.; Yan, W. Prestress force identification for externally pre–stressed concrete beam based on frequency equation and measured frequencies. Math. Probl. Eng. 2014, 2014, 840937. [Google Scholar] [CrossRef]
  63. Xiang, Z.; Chan, T.H.T.; Thambiratnam, D.P.; Nguyen, A. Prestress and excitation force identification in a prestressed concrete box-girder bridge. Comput. Concr. 2017, 20, 617–625. [Google Scholar]
  64. Breccolotti, M. On the evaluation of prestress loss in PRC beams by means of dynamic techniques. Int. J. Concr. Struct. Mater. 2018, 12, 1. [Google Scholar] [CrossRef] [Green Version]
  65. Breccolotti, M.; Pozza, F. Prestress evaluation in continuous PSC bridges by dynamic identification. Struct. Monit. Maint. 2018, 5, 463–488. [Google Scholar]
  66. Law, S.S.; Wu, S.Q.; Shi, Z.Y. Moving load and prestress identification using wavelet based method. J. Appl. Mech. Trans. ASME 2008, 75, 021014. [Google Scholar] [CrossRef]
  67. Li, H.; Lv, Z.; Liu, J. Assessment of prestress force in bridges using structural dynamic responses under moving vehicles. Math. Probl. Eng. 2013, 2013, 435939. [Google Scholar] [CrossRef]
  68. Xiang, Z.; Chan, T.H.T.; Thambiratnam, D.P.; Nguyen, T. Synergic identification of prestress force and moving load on prestressed concrete beam based on virtual distortion method. Smart Struct. Syst. 2016, 17, 917–933. [Google Scholar] [CrossRef]
  69. Tullini, N.; Laudiero, F. Dynamic identification of beam axial loads using one flexural mode shape. J. Sound Vib. 2008, 318, 131–147. [Google Scholar] [CrossRef]
  70. Bahra, A.S.; Greening, P.D. Identifying multiple axial load patterns using measured vibration data. J. Sound Vib. 2011, 330, 3591–3605. [Google Scholar] [CrossRef]
  71. Tullini, N.; Rebecchi, G.; Laudiero, F. Bending tests to estimate the axial force in tie–rods. Mech. Res. Commun. 2012, 44, 57–64. [Google Scholar] [CrossRef]
  72. Maes, K.; Peeters, J.; Reynders, E.; Lombaert, G.; De Roeck, G. Identification of axial forces in beam members by local vibration measurements. J. Sound Vib. 2013, 332, 5417–5432. [Google Scholar] [CrossRef] [Green Version]
  73. Rebecchi, G.; Tullini, N.; Laudiero, F. Estimate of the axial force in slender beams with unknown boundary conditions using one flexural mode shape. J. Sound Vib. 2013, 332, 4122–4135. [Google Scholar] [CrossRef]
  74. Tullini, N. Bending tests to estimate the axial force in slender beams with unknown boundary conditions. Mech. Res. Commun. 2013, 53, 15–23. [Google Scholar] [CrossRef]
  75. Luong, H.T.M.; Zabel, V.; Lorenz, W.; Rohrmann, R.G. Vibration-based model updating and identification of multiple axial forces in truss structures. Procedia Eng. 2017, 188, 385–392. [Google Scholar] [CrossRef]
  76. Luong, H.T.M.; Zabel, V.; Lorenz, W.; Rohrmann, R.G. Non-destructive assessment of the axial stress state in iron and steel truss structures by dynamic measurements. Procedia Eng. 2017, 199, 3380–3385. [Google Scholar] [CrossRef]
  77. Bonopera, M.; Chang, K.C.; Chen, C.C.; Lee, Z.K.; Tullini, N. Axial load detection in compressed steel beams using FBG–DSM sensors. Smart. Struct. Syst. 2018, 21, 53–64. [Google Scholar]
  78. Bonopera, M.; Chang, K.C.; Chen, C.C.; Lin, T.K.; Tullini, N. Compressive column load identification in steel space frames using second–order deflection–based methods. Int. J. Struct. Stab. Dyn. 2018, 18, 1850092. [Google Scholar] [CrossRef]
  79. Bonopera, M.; Chang, K.C.; Chen, C.C.; Lin, T.K.; Tullini, N. Bending tests for the structural safety assessment of space truss members. Int. J. Space Struct. 2018, 33, 138–149. [Google Scholar] [CrossRef]
  80. Bonopera, M.; Chang, K.C.; Chen, C.C.; Sung, Y.C.; Tullini, N. Feasibility study of prestress force prediction for concrete beams using second–order deflections. Int. J. Struct. Stab. Dyn. 2018, 18, 1850124. [Google Scholar] [CrossRef]
  81. Bonopera, M.; Chang, K.C.; Chen, C.C.; Sung, Y.C.; Tullini, N. Prestress force effect on fundamental frequency and deflection shape of PCI beams. Struct. Eng. Mech. 2018, 67, 255–265. [Google Scholar]
  82. Kernicky, T.; Whelan, M.; Al-Shaer, E. Dynamic identification of axial force and boundary restraints in tie rods and cables with uncertainty quantification using Set Inversion Via Interval Analysis. J. Sound Vib. 2018, 423, 401–420. [Google Scholar] [CrossRef]
  83. Noble, D.; Nogal, M.; O’Connor, A.; Pakrashi, V. Dynamic impact testing on post–tensioned steel rectangular hollow sections; An investigation into the “compression–softening” effect. J. Sound Vib. 2015, 355, 246–263. [Google Scholar] [CrossRef] [Green Version]
  84. Tullini, N.; Rebecchi, G.; Laudiero, F. Reliability of the tensile force identification in ancient tie-rods using one flexural mode shape. Int. J. Archit. Herit. 2019, 13, 402–410. [Google Scholar] [CrossRef]
  85. Turco, E. Identification of axial forces on statically indeterminate pin-jointed trusses by a nondestructive mechanical test. Open Civ. Eng. J. 2013, 7, 50–57. [Google Scholar] [CrossRef] [Green Version]
  86. Hsiao, J.K. Prestress loss distributions along simply supported pretensioned concrete beams. Electron. J. Struct. Eng. 2017, 16, 18–25. [Google Scholar]
  87. ASTM (American Society for Testing and Materials). Annual Book of ASTM standards, Section 4: Construction vol. 04.02; Concrete & Aggregates: West Conshohocken, PA, USA, 2016. [Google Scholar]
  88. Chiu, Y.T.; Lin, T.K.; Hung, H.H.; Sung, Y.C.; Chang, K.C. Integration of in–situ load experiments and numerical modeling in a long–term bridge monitoring system on a newly–constructed widened section of freeway in Taiwan. Smart Struct. Syst. 2014, 13, 1015–1039. [Google Scholar] [CrossRef]
  89. Sung, Y.C.; Lin, T.K.; Chiu, Y.T.; Chang, K.C.; Chen, K.L.; Chang, C.C. A bridge safety monitoring system for prestressed composite box–girder bridges with corrugated steel webs based on in–situ loading experiments and a long–term monitoring database. Eng. Struct. 2016, 126, 571–585. [Google Scholar] [CrossRef]
  90. Jeon, J.C.; Lee, H.H. Development of displacement estimation method of girder bridges using measured strain signal induced by vehicular loads. Eng. Struct. 2019, 186, 203–215. [Google Scholar] [CrossRef]
  91. Bonopera, M.; Chang, K.C.; Chen, C.C.; Lee, Z.K.; Sung, Y.C.; Tullini, N. Fiber Bragg grating–differential settlement measurement system for bridge displacement monitoring: Case study. J. Bridge Eng. 2019, 24, 05019011. [Google Scholar] [CrossRef]
Figure 1. Static NDT methods proposed by Tullini et al. [71,74]: (a) cast iron disks suspended at the reference beam; (b) arrangement of the dial indicators for vertical deflection measurements during an experimental simulation. Copyright © 2012, 2013 Elsevier Ltd. Reprinted with permission.
Figure 1. Static NDT methods proposed by Tullini et al. [71,74]: (a) cast iron disks suspended at the reference beam; (b) arrangement of the dial indicators for vertical deflection measurements during an experimental simulation. Copyright © 2012, 2013 Elsevier Ltd. Reprinted with permission.
Applsci 10 07257 g001
Figure 2. Reference model of the PC girder [80]: (a) deflection curve v(0) after the application of the eccentric prestress force N; (b) deflection curve v(1) after the application of the vertical load F to the deflection curve v(0). The dashed lines represent the initial deflection curves. Copyright © 2018, World Scientific Publishing Co. Pte Ltd. Reprinted with permission.
Figure 2. Reference model of the PC girder [80]: (a) deflection curve v(0) after the application of the eccentric prestress force N; (b) deflection curve v(1) after the application of the vertical load F to the deflection curve v(0). The dashed lines represent the initial deflection curves. Copyright © 2018, World Scientific Publishing Co. Pte Ltd. Reprinted with permission.
Applsci 10 07257 g002
Figure 3. Three-point bending tests on a PCI girder conducted in the laboratory of NCREE, Taiwan [50]: (a) indoor test rig; (b) application of a FBG-Differential Settlement Measurement (DSM) liquid-level system for vertical deflection measurements along the span. Copyright © 2019 Elsevier Ltd. Adapted with permission.
Figure 3. Three-point bending tests on a PCI girder conducted in the laboratory of NCREE, Taiwan [50]: (a) indoor test rig; (b) application of a FBG-Differential Settlement Measurement (DSM) liquid-level system for vertical deflection measurements along the span. Copyright © 2019 Elsevier Ltd. Adapted with permission.
Applsci 10 07257 g003
Figure 4. Prestress force determinations Na based on (a) Equation (8a) (Test 1) and (b) Equation (8b) (Test 2). Symbols + refer to the comparison between determined Na and measured values of prestress force N for all the 27 test combinations. The dashed lines with symbol × represent the sensitivity analyses.
Figure 4. Prestress force determinations Na based on (a) Equation (8a) (Test 1) and (b) Equation (8b) (Test 2). Symbols + refer to the comparison between determined Na and measured values of prestress force N for all the 27 test combinations. The dashed lines with symbol × represent the sensitivity analyses.
Applsci 10 07257 g004
Table 1. Laboratory investigations on the relation between prestress force and dynamics of Prestressed Concrete (PC) girders.
Table 1. Laboratory investigations on the relation between prestress force and dynamics of Prestressed Concrete (PC) girders.
AuthorYearVibration TestResult and Finding
Hop [31]1991UnCracked PC girdersIncrease in prestress force slightly increases the natural frequencies
Saiidi et al. [32]1994UnCracked PC girder(as above)
Miyamoto et al. [33]2000UnCracked PC girders(as above)
Lu and Law [34]2006UnCracked PC girder(as above)
Xiong and Zhang [35]2009Cracked PC girdersIncrease in prestress force slightly decreases the natural frequencies
Kim et al. [36]2010Cracked PC girderIncrease in prestress force slightly increases the natural frequencies
Jang et al. [37]2010UnCracked PC girders(as above)
Noh et al. [38]2015UnCracked PC girders(as above)
Table 2. Numerical investigations on the relation between prestress force and dynamics of PC girders.
Table 2. Numerical investigations on the relation between prestress force and dynamics of PC girders.
AuthorYearNumerical SolutionDynamic ModelResult and Finding
Chan and Yung [39]2000AnalyticalUnCracked PC girdersIncrease in prestress force slightly decreases the natural frequencies
Kim et al. [43]2004Analytical simulation of tests performed by Saiidi et al. [32]UnCracked PC girdersIncrease in prestress force slightly increases the natural frequencies
Law and Lu [44]2005AnalyticalUnCracked PC girdersIncrease in prestress force slightly decreases the natural frequencies
Hamed and Frostig [45]2006AnalyticalUnCracked PC girdersIncrease in prestress force does not affect the natural frequencies
Jaiswal [46]2008FEUnCracked PC girders(as above)
Limongelli et al. [47]2016AnalyticalUnCracked and Cracked PC girders(as above)
Gan et al. [48]2019FE simulation of tests performed by Jang et al. [37] and Noble et al. [49]Cracked PC girdersIncrease in prestress force slightly increases the natural frequencies
Bonopera et al. [50]2019AnalyticalUnCracked PC girdersIncrease in prestress force does not affect the natural frequencies
Luna Vera et al. [53]2020AnalyticalUnCracked and Cracked PC girders(as above)
Bonopera et al. [51]2021AnalyticalUnCracked PC girders(as above)
Table 3. Testing methods for determining prestress losses in PC girders.
Table 3. Testing methods for determining prestress losses in PC girders.
MethodReferenceTesting on a PC GirderResult and Finding
(1)
Destructive
[23,28,54,55]Static test to determine crack initiation or crack re-opening loads to obtain the compressive stress in the bottom flangeIt causes damages
(2)
Destructive
[23,56]Severing the prestressing tendon by cutting it into an exposed length after placing strain gaugesIt causes damages
(3)
Semidestructive
[54,57]Relating the tension in the tendon to a vertical deflection recorded when weights are suspended from it on an exposed lengthIt causes partial damages
(4)
Nondestructive
[55,58]Determining the side pressure to close the induced crack in a small cylindrical hole drilled adjacent to the tendon in the bottom flangeThe embedded tendon, if is too close to the concrete surface, could be damaged by the deep hole-cut
(5)
Nondestructive
[59,60,61,62,63,64,65,66,67,68]Vibration test to determine natural frequencies and/or dynamic responsesIt requires an accurate selection of the mode shape
Table 4. Prestress force determinations Na based on Equation (8a) (Test 1) and Equation (8b) (Test 2), and corresponding measured parameters for each test day [80]. Copyright © 2018, World Scientific Publishing Co. Pte Ltd. Adapted with permission.
Table 4. Prestress force determinations Na based on Equation (8a) (Test 1) and Equation (8b) (Test 2), and corresponding measured parameters for each test day [80]. Copyright © 2018, World Scientific Publishing Co. Pte Ltd. Adapted with permission.
Test 1-v2(1)
Deflections at a Quarter
Test 2-v4(1)
Deflections at the Midspan
Days of Concrete CuringEaverNFNaΔNaΔ
(MPa)(kN)(kN)(kN)(%)(kN)(%)
42634,87062020.277725.378927.3
62022.685738.285738.2
61725.0386−37.4550−10.9
42737,61872420.17290.77321.1
72122.67210.07615.5
72125.1715−0.8718−0.4
43338,79182020.28463.28230.4
82022.98250.68250.6
82025.18240.58706.1
Table 5. Characteristics of the static NDT method preliminary proposed by Bonopera et al. [80] for determining prestress losses in PC girders.
Table 5. Characteristics of the static NDT method preliminary proposed by Bonopera et al. [80] for determining prestress losses in PC girders.
AdvantagesDisadvantages
(1) Precise determinations by vertical deflection measurements of 0.01 mm in accuracy(1) Determination of the concrete elastic modulus by compression tests on a set of drilled cores
(2) No requirement of direct measure of the tension force in the tendon(2) Vertical deflections, of 0.01 mm in accuracy, are not always easy to measure in situ
(3) No requirement of any selecting experimental data
(4) Determinations take into account the combined effects of tendon relaxation, concrete creep and shrinkage, and parameters of the real environment
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bonopera, M.; Chang, K.-C.; Lee, Z.-K. State-of-the-Art Review on Determining Prestress Losses in Prestressed Concrete Girders. Appl. Sci. 2020, 10, 7257. https://doi.org/10.3390/app10207257

AMA Style

Bonopera M, Chang K-C, Lee Z-K. State-of-the-Art Review on Determining Prestress Losses in Prestressed Concrete Girders. Applied Sciences. 2020; 10(20):7257. https://doi.org/10.3390/app10207257

Chicago/Turabian Style

Bonopera, Marco, Kuo-Chun Chang, and Zheng-Kuan Lee. 2020. "State-of-the-Art Review on Determining Prestress Losses in Prestressed Concrete Girders" Applied Sciences 10, no. 20: 7257. https://doi.org/10.3390/app10207257

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