Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Received: 5 February 2018 Revised: 18 December 2018 Accepted: 29 March 2019

DOI: 10.1002/stc.2374

REVIEW

Monitoring of prestressing forces in prestressed concrete


structures—An overview

Hiba Abdel-Jaber Branko Glisic

Department of Civil and Environmental


Engineering, Princeton University, Summary
Princeton, New Jersey
This paper presents an overview of currently available methods for monitoring
Correspondence prestressing forces in prestressed concrete structure. Structural health moni-
Hiba Abdel-Jaber, Department of Civil toring has become an increasingly important tool for assessment of structural
and Environmental Engineering,
performance. Additionally, the value of the prestressing force represents an
Princeton University, Princeton, NJ 08544.
Email: [email protected] important parameter in prestressed concrete structures. Thus, several methods
for monitoring prestress forces have emerged. This paper aims to consolidate
Funding information
Division of Civil, Mechanical and
the work performed in the area of prestressing force monitoring by present-
Manufacturing Innovation, Grant/Award ing the most important advances and the directions for future research. The
Number: 1362723; U.S. Department of methods presented in this paper are based on indirect monitoring of the pre-
Transportation, Grant/Award Number:
DTRT13-G-UTC28 stressing force through monitoring of another relevant parameter. They are
divided into five general classes based on the relevant parameter they monitor:
(a) vibration-based methods (based on acceleration), (b) impedance-based meth-
ods (based on electrical impedance), (c) acoustoelastic methods (based on wave
velocity), (d) elasto-magnetic methods (based on magnetic permeability), and (e)
strain-based methods (based on strain). The paper presents a table summarizing
the comparison between the methods based on defined criteria.

K E Y WO R D S
force monitoring, prestressed concrete, prestress loss, structural health monitoring

1 I N T RO DU CT ION
Since the introduction of prestressed concrete in the 1940s, there has been a significant increase in its use as a construction
material due to its numerous advantages over traditional reinforced concrete, such as reduced member sizes, deflection
control, and cracking control. For example, more than 45% of bridges built in the United States since 2010 have been
made of prestressed concrete.1 Additionally, prestressed concrete is routinely used in the construction of several other
types of important structures such as containment vessels of nuclear reactors, silos and tanks, posttensioned floor slabs,
and prestressed and posttensioned concrete wind turbine towers. Due to the importance of these structures, extending
their service life is of interest. This has led to several efforts in the monitoring of prestressed concrete to ensure structural
integrity and performance through both long-term structural health monitoring (SHM) and intermittent nondestructive
evaluation (NDE).
Prestress force levels are indicative of damage as both a precursor to certain types of damage (their loss can cause cracks)
and a consequence of other types of damage (corrosion or rupture of strands). If force levels drop below a certain thresh-
...............................................................................................................................................................
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the
original work is properly cited.
© 2019 The Authors. Structural Control and Health Monitoring Published by John Wiley & Sons Ltd

Struct Control Health Monit. 2019;e2374. wileyonlinelibrary.com/journal/stc 1 of 27


https://doi.org/10.1002/stc.2374
2 of 27 ABDEL-JABER AND GLISIC

old, tensile stresses can develop in the concrete, resulting in cracking or excessive deflections. Additionally, corrosion or
rupture in prestressing strands result in local changes in the prestressing force levels, which can have a catastrophic effect
on the structure. If a local change in prestressing force levels is detected, it can be indicative of such types of damage
and maintenance can be scheduled to prevent propagation of this damage to a level that can threaten the safety of the
structure.
Prestressing force levels are expected to decrease over the service life of a structure due to time-dependent losses that
are caused by steel strand relaxation and concrete dimensional changes due to creep and shrinkage. These prestress losses
are accounted for in the design process using empirical equations from design codes. However, these empirical equations
are typically based on laboratory tests and may not truly represent the actual concrete mix and curing conditions. For
example, prestress losses in prestressed high performance concrete have been shown to be significantly different from
design predictions and dependent on the concrete mix.2-6 Overestimating prestress losses may result in conservative and
possibly uneconomic designs. On the other hand, underestimating prestress losses can result in tensile stresses in the
concrete and possibly cracking, which can affect durability by exposing reinforcement to the environment, and in some
cases, combined with other mechanisms (e.g., corrosion), can lead to failure. Thus, evaluation of prestress forces and thus
losses in prestressed concrete structures has become increasingly important.
The aim of this paper is to present and discuss the state of the art in monitoring prestressing force levels in prestressed
concrete in laboratory and field applications. In particular, the aim is to summarize the progress made thus far in each of
the fields of methods of monitoring of prestressing force and to identify their advantages and limitations with regard to
specific application. Finally, this paper aims to identify challenges to the application of methods on real structures and to
make recommendations to overcome said challenges.
This paper includes a representative but not exhaustive literature review due to the density of studies in the field of
monitoring prestressing forces. To the best of the authors' knowledge, the paper represents all active research areas in
prestress force monitoring. Because the purpose is to present the state of the art of the field, and in the interest of brevity,
literature included in this paper represents the progress milestone studies in each area in order to provide a complete
perspective rather than a summary of every study performed.
Currently, directly measuring force levels in a structure in field applications is not feasible. Although forces can be mon-
itored by means of a hydraulic force gauge, this would only provide information of the force near the anchorage (where
the senor is usually installed). Additionally, no literature has been found that demonstrates the durability and longevity
of such force sensors in field applications. Instead, all technologies available and presented here rely on measuring a
parameter that can be correlated with force levels. The paper is organized into five main sections, each of which discusses
methods based on one parameter or a group of similar parameters. The methods presented are as follows: vibration meth-
ods (based on acceleration), impedance-based methods (based on electrical impedance), acoustoelastic methods (based
on wave velocity), elasto-magnetic (EM) methods (based on magnetic permeability), and strain-based methods (based on
strain). Each method is presented with considerable detail so that a less experienced reader can get sufficient informa-
tion; however, a more experienced reader who might be less interested in details can move directly to a brief summary on
the method included at the end of each section.

2 EVALUATION CRITERIA
The methods presented in this paper consider general classes of NDE and SHM methods and not specific sensors or
algorithms. Thus, they are evaluated based on their overall promise to provide stable and accurate long-term information
on prestressing forces. The five criteria for evaluation include sensitivity to prestress force changes, imperviousness to
other regularly changing factors such as environmental variables, ability to monitor the local distribution of the force along
the strand, feasibility of instrumentation, and ease of applicability to different types of real-life structures. It should be
noted that this paper examines methods pertaining mostly to strands in prestressed concrete structures and cable-stayed
structures.
An ideal method should be highly sensitive to minor changes in prestressing force, while being completely unaffected
by changes in other factors such as temperature and humidity, which are expected to vary periodically throughout the
life of the structure. Prestress losses are typically less than 25% of the initial prestress force value, and the average based
on more than forty studies was reported to be 21%.7 Although the variation of the prestress loss within several percent
in general is not a concern for design as long as the loss stabilizes over time, steady evolution or sudden increase of the
losses over time can be critical. In these cases, sensitivity of the method is very important because methods with higher
sensitivity are more likely to detect evolution or increase of prestress losses at an early stage, which helps managers to
plan preventive or repair actions.
ABDEL-JABER AND GLISIC 3 of 27

Imperviousness to environmental factors, particularly temperature, is among the major challenges to on-site imple-
mentations of most long-term SHM and NDE techniques, regardless of the purpose of monitoring. Because no monitoring
method or parameter is completely unaffected by temperature changes, the criterion in this case relates to the degree to
which the influence of temperature on collected measurements, and thus results, is known and understood.
The force distribution along the strand provides important information regarding defects in the strand, which have been
reported in the literature,8,9 as well as regarding immediate losses in prestressed concrete, such as those due to anchorage
set and friction. Significant change in the magnitude of the force along the strand can imply wire breakage, the effect of
which is negligible as the distance from the breakage increases, and might be undetected if the force is only monitored
near the anchorage. This criterion relates to the feasibility of instrumenting the strand or beam structure with multiple
sensors along the longitudinal direction. Additionally, it relates to the localness of the monitored parameter, that is, the
region in which it is valid and representative. For example, natural frequency is a global parameter that characterizes the
entire structure, whereas strain is a local parameter representative of the behavior at the particular location at which it is
measured.
The feasibility and potential for on-site implementations is another important criterion in the evaluation and com-
parison of the methods. This criterion is affected by the cost, size, and weight of sensors and equipment required for a
monitoring technique. Thus, it strongly relates to the maturity of the monitoring technique. More mature techniques
have more readily available commercial sensors of various sizes depending on the purpose of instrumentation, thereby
reducing the cost of monitoring and providing good quality control of sensors, as opposed to those in the research phase.
Applicability to different types of structures is intrinsic to the method and is controlled by two aspects: (a) requirement
for direct access to the prestressing strand and (b) need for a finite element model for data analysis. Both of these aspects
limit the types of structures that can be monitored using a given method in different ways; the requirement for direct access
limits the use of the method to structures with external prestressing tendons, whereas the need for a finite element model
limits the application to less complex structures with known material properties. While overcoming the requirement
for direct access is difficult, overcoming the need for a finite element model is often possible through calibration with
known prestress force levels, which allows for creating an empirical relationship between the measured parameter and the
prestress force level without the need for a finite element model. However, site conditions and requirements do not always
allow for a calibration phase with known force levels, rendering this option more appropriate for laboratory applications.

3 VIBRATION METHODS

Vibration methods encompass all methods of prestressing force identification that utilize the vibration properties of a
structure. For the purposes of this study, presented methods will be limited to those involving natural frequencies as
modal shapes or displacement measurements were shown to be unaffected by change in prestressing forces.10 The main
principle is based on the dependence of natural frequencies on the stiffness of the structure, which in turn is debatably
dependent on the prestressing force level. Such methods have been proven successful for cable-stayed structures where
the force in the cable is directly related to the measured natural frequency using the tout-string theory.11-13 However, the
relationship is more complicated in prestressed concrete structures. Although many researchers have developed relation-
ships between the prestressing force and resulting natural frequencies of the structure, other researchers have shown
that the prestressing force has no effect on natural frequencies. There are three different arguments for the relationship
between prestressing forces and natural frequency.
First, some researchers argue that an increase in prestressing forces results in a decrease in natural frequencies, mainly
due to the effect of “compression softening.”14,15 One of the first attempts to relate the vibration frequency of a structure
to the prestress force level was an experimental study by Saiidi et al.14 The study first examined an analytical model for
the relationship, given by Equation 1, which theorized that increasing prestress force levels would cause a decrease in
natural frequencies.
( )2 ( )4
n𝜋 N n𝜋 EI
𝜔2n = − + , (1)
L m L m
where 𝜔 is the natural frequency of vibration, n is the mode number, L is the span length, N is the axial compressive force,
that is, prestressing force, m is the beam mass per unit length, E is the modulus of elasticity of the concrete beam, and I
is the moment of inertia for the beam section.
However, laboratory and field tests carried out in the study showed the opposite effect, where increasing prestress
force levels caused increased natural frequencies,14 a result that was experimentally confirmed in other studies.16 Thus,
4 of 27 ABDEL-JABER AND GLISIC

the second argument, mainly proven through experimental results, claims that an increase in prestress force causes an
increase in natural frequencies. The discrepancy was explained by the stiffening effect due to closing of microcracks in
the concrete as prestress force levels increase, thereby increasing the stiffness and natural frequencies of the structure, an
effect not taken into account in the analytical model, given by Equation 1. The authors confirmed this effect by considering
the change in effective flexural rigidity (i.e., flexural rigidity that would result in the measured natural frequencies) with
the change in axial force. The positive correlation is presented in Figure 1.
Other researchers agree that increasing prestress forces in cracked beams results in crack closure that increases stiffness.
Noble et al17 compared the natural frequencies of an uncracked and a cracked beam with varying levels of prestressing
force. As shown in Figure 2, the study showed no statistically significant relationship between prestress force and natural
frequencies for uncracked beam (Figure 2a) but a significant relationship for a cracked beam (Figure 2b). The researchers
further explained the complicated relationship in Figure 2b, where there is a significant decrease in natural frequencies
for prestressing force levels under 160 kN, followed by an increasing trend. The researchers attributed the decrease to
the complex dynamic behavior of the cracked beam causing it to behave as a series of independent beams with lengths
that increase as closure of cracks occurs. The increasing span lengths cause a decrease in the stiffness and thus natural
frequencies of the structure. However, as cracks continue to gradually close due to the increasing prestressing force, the

FIGURE 1 Variation of flexural rigidity with axial (prestressing) force14

FIGURE 2 Regression analysis of the fundamental frequency with posttensioning load for (a) an uncracked beam and (b) a cracked beam17
ABDEL-JABER AND GLISIC 5 of 27

beam begins to behave monolithically at a certain force threshold (160 kN in this experiment), resulting in an increase in
stiffness and natural frequencies as prestressing force increases.
Other researchers have reported a positive correlation between crack closure and natural frequencies, further contribut-
ing to the second argument. Grace and Ross experimentally showed a decrease in natural frequencies of posttensioned
concrete beams caused by a decrease in stiffness due to fatigue cracking.18 Hop experimentally showed similar results.19
Hamed and Frostig analytically showed a reduction in natural frequencies with increased cracking, but no model for
the relationship to prestress force levels was developed.20 De Roeck agrees with the researchers and argues that loss of
prestress can only be detected by vibration methods if accompanied by cracking.21
The discrepancy gave rise to the third argument that, if cracking is neglected, prestressing forces do not affect natural
frequencies.22-24 Dall'Asta and Dezi present a simplified but accurate analytical model for the natural frequencies of a
prestressed concrete beam that accounts for the fact that the prestressing force is internal to the system, given by Equation
2 (as opposed to Equation 1 that assumes the prestressing force to be external).22
[( ) ( ) ]
n4 𝜋 4 N N
𝜔n =
2
Eb − Ib + E c + Ic (2)
mL4 Ab Ac
where Eb , Ab , and Ib are the elastic modulus, cross-sectional area, and moment of inertia of the concrete beam, respectively,
and Ec , Ac , and Ic are the same quantities for the steel strand.
Based on typical values for a prestressed concrete beam, the terms N∕Ab and N∕Ac are negligible compared to the terms
Eb and Ec , respectively. Furthermore, Ic is negligible compared to Ib , reducing the equation to that of a simply supported
beam and implying that prestressing forces do not have a significant impact on natural frequencies of the beam.
The other two discussions of the study by Saiidi et al similarly support that prestressing forces should not be treated as
an external force in the system and thus do not cause compression softening or a frequency change, except due to closing
of microcracks.23,24 This was confirmed by another study where researchers compared the effect of posttensioning steel
hollow beams and externally axially loading them to test the theory of compression softening without the interference of
microcracks.25 The conclusion was that posttensioning and external axial loads do not affect beams in the same manner,
with only the latter causing compression softening in slender beams. Sample results from the study are presented in
Figure 3. The labels Beam 1 and Beam 2 refer to a slender beam and a stocky beam, respectively, to test the theory that
compression softening affects beams susceptible to buckling (slender beam only), whereas the labels Case 1 and Case
2 refer to an externally axially loaded beam and a posttensioned beam, respectively. As shown in Figure 3a, in the case
of a slender beam, increasing external axial load causes behavior that complies with the compression softening theory,
whereas increasing prestress force causes a statistically significant decrease in the fundamental bending frequency, but
that behavior does not comply with the compression softening theory. In Figure 3b, neither the externally axially loaded
nor the posttensioned beam show behavior that agrees with compression softening theory; an externally axially loaded
beam shows a significant increase in bending frequency, whereas a posttensioned beam shows a statistically significant
decrease in bending frequency with increasing force.
Furthermore, a rigorous analytical model derived by Hamed and Frostig26 using the variational principle of virtual work
shows that there is no dependence of natural frequencies on prestressing forces. Moreover, another extensive experimental

FIGURE 3 Means of dynamic test data with varying load for (a) a slender beam and (b) a stocky beam25
6 of 27 ABDEL-JABER AND GLISIC

program on uncracked prestressed concrete beams showed that there is no statistically significant change in natural fre-
quencies due to change in prestress force for six of the nine tested beams,27 with the authors additionally arguing that any
observed changes are due to chance and not systematic change.
The contradicting arguments show that prestress force identification based on vibration methods has not been achieved.
However, the following conclusions can be drawn from the presented studies:
1. The compression softening theory for prestressed concrete beams has been debunked based on analytical work, as
given in Equation 2 and experimental studies, as shown in Figure 3.
2. In the case of uncracked beams, changes in prestress forces have no effect on natural frequencies, as shown in
Figure 2a and discussed in other studies.
3. In the case of cracked beams, an increase in prestress forces causes closure of microcracks, which stiffens the struc-
ture and causes an increase in natural frequencies. The increase in rigidity has been reported in multiple studies as
shown in Figures 1 and 2b.
Based on the studies and above conclusions, detection of prestress loss using changes in natural frequencies is not possi-
ble in uncracked beams. Additionally, even in the case of cracked beams, there is no developed relation between the level
of prestressing and the natural frequencies because it is dependent on the level of cracking. This is further complicated by
the global nature of vibration monitoring because other types of damage and change of boundary conditions can result
in changes in natural frequencies. Moreover, natural frequencies can vary significantly due to environmental changes28
and lack of ideal supports; one study reports a 6% change,29 whereas another reports up to 18% change.30 In comparison,
changes in frequency due to prestress losses are insignificant and may not be detected.27

4 IMPEDANCE-BASED METHODS

Impedance-based SHM methods exploit properties of piezoelectric materials for SHM purposes. In piezoelectric materials,
electrical and mechanical responses are coupled; mechanical stress produces electrical charge and vice versa.31 This is
known as the piezoelectric effect and both the direct and converse versions are used in sensing applications, thereby
allowing a piezoelectric material to function as both an actuator and sensor, typically by means of a piezoceramic material,
such as PZT (Lead Zirconate Titanate). A PZT patch is affixed to a structural element and an alternating electrical field
is applied to it. Due to the converse piezoelectric effect, a deformation is induced in the PZT patch and consequently the
host structure. This engages the direct piezoelectric effect, and the induced deformation causes an electric charge. Liang
et al.32 first showed that the resulting electrical impedance of the PZT is related to the mechanical impedance of the host
structure. As given in Equation 3, the electrical admittance, Y(𝜔), (inverse of electrical impedance) of the PZT actuator is
a function of the mechanical impedance of both the PZT actuator, Z(𝜔), and the host structure, Za (𝜔).
( )
I Z(𝜔) 2 ̂E
Y (𝜔) = = i𝜔a 𝜀−T − d Y , (3)
V 33
Z(𝜔) + Za (𝜔) 3x xx
where V is the input voltage to the PZT actuator, I is the output current from the PZT sensor, and a, d23x , Ŷ xx
E
, and 𝜀−T
33
are
the geometry constant, piezoelectric coupling constant, Young's modulus, and the complex dielectric constant of the PZT
at zero stress, respectively.
Assuming the mechanical properties of the PZT sensor do not change, changes in the electrical impedance signature
reflect changes in the mechanical properties of the host structure, thereby allowing PZT sensors to be used for damage
detection purposes. Due to their small size and negligible effects on the properties and behavior of the structure, as shown
in Figure 4b, PZT sensors offer an advantage over other sensing technologies.
Several applications of PZT impedance-based SHM have emerged including detection of cracks35-37 and
delamination.38,39 The first attempt to monitor prestressed concrete structures by PZT sensors was carried out by Abe
et al,40 where the researchers developed the analytical relationship between stress and the wave propagation (directly
related to electrical admittance) for both the one-dimensional case (bar under axial stress) and two-dimensional case
(plate under axial stress). Based on measured impedance and the analytical relationship, the absolute stress state of the
elements was estimated,40 but the estimates were lower than measurements made by strain gauges due to stress transfer
at the interface of the PZT and the structural element.
Subsequent studies focused on the detection of changes in the prestressing force, rather than the absolute force or stress
due to the difficulty of analytically relating the stress to the PZT measurements, as illustrated by the aforementioned
ABDEL-JABER AND GLISIC 7 of 27

FIGURE 4 Instrumentation of prestressed concrete beam used in the studies of Kim et al33,34 showing piezoelectric sensor in the form of
Macro-Fiber Composite (MFC) patch on anchor plate, highlighted with an arrow

FIGURE 5 Changes in impedance signature with (a) decreasing prestress force levels and (b) increasing stiffness near the center of the
beam34

study. Kim et al33 combined a global vibration-based method with a local impedance-based method to achieve three-step
SHM: (a) damage detection using a damage index based on global frequency responses, that is, detection of prestress loss,
(b) characterization of prestress loss using impedance signals, that is, confirming that the change to the damage index is
due to prestress loss and not changes to boundary conditions, and (c) prestress loss estimation using modal parameters.
The method was applied to a 6-m prestressed concrete girder in the laboratory instrumented with seven equally spaced
accelerometers along the span, as well as a load cell on the jacking end and a PZT patch on the dead end. Five different
levels of prestress force were applied to the beam, simulating four prestress loss stages, each of which constituted 17%
loss of the initial prestressing force. Prestress loss was characterized using the frequency shift in the impedance signal,
shown in Figure 5a, and quantified using changes in natural frequencies. As shown in Figure 5, as prestressing force levels
were decreased, the peak frequency of the impedance signal shifted to the right. The sensitivity and accurate prediction
of smaller changes in the prestressing force value were not demonstrated.
A subsequent study extended the algorithm to additionally identify a second type of damage, added mass to the beam
to simulate stiffness change.34 It was shown that impedance signatures did not significantly change due to stiffness
change, as shown in Figure 5b, indicating that the method is robust to other changes to the structure that do not impact
the prestress force. Despite a demonstration of the capability of impedance-based methods to characterize prestress loss
damage demonstrated in the previous two studies, the studies feature two limitations: (a) the necessity of the use of vibra-
tion characteristics for quantification of prestress loss, which is not a simple task as discussed in the previous section
(vibration-based methods) and (b) the requirement for high resonant frequency range to establish sensitivity to changes
in prestressing forces. In both of the studies, the frequency range required for achieving impedance signatures sensitive
to prestress loss change was determined through trial and error to be 880–980 kHz, as shown in Figure 5a. Such a high
8 of 27 ABDEL-JABER AND GLISIC

frequency range is disadvantageous for two reasons: (a) It requires the use of a high performance impedance analyzer
that is both heavy and expensive, making it unsuitable for typical site deployment, and (b) it does not allow for wireless
deployment because the frequency range of wireless sensors is less than 100 kHz.41
To lower the frequency range, researchers investigated the possibility of installing an aluminum plate between the
bearing plate and the anchor block and affixing the PZT patch to the plate,41,42 as shown in Figure 6. Because the stiff-
ness of the aluminum plate is lower than that of the bearing plate and the anchor block, the resonant frequency range
is reduced. Additionally, Nguyen and Kim42 extended the study to analytically determine the frequency range before the
experiment to overcome the difficulty of the trial-and-error method in field applications, which is often not feasible. In
both applications, laboratory experiments were performed and the change in impedance due to changes in prestress force
was detected using wireless sensors with significant frequency range reduction compared to previous studies (20–30 kHz
in this case), as shown in Figure 7a.41,42 Empirical relationships were derived between prestress loss and frequency shift,
shown in Figure 7b, as well as between prestress loss and a damage index. However, the relationships are only applica-
ble to the structure and environmental conditions for which they were derived and cannot be generalized.42 It is worth
noting that because PZT sensors measure local effects, prestress loss levels detected and quantified reflect changes at the
anchorage only and not necessarily throughout the length of the beam structure.
Huynh and Kim43 recognized that the installation of a plate between the bearing plate and the anchor block reduces
the stiffness of the connection due to the low stiffness of the plate and also limits the application of the method to new
structures. To extend the method to existing structures, they designed a portable PZT interface consisting of a thin-walled

FIGURE 6 Setup of the PZT patch attached to an aluminum plate (interface washer): (a) schematic setup and (b) installed at anchorage of
prestressed concrete beam41

FIGURE 7 Detection of prestress loss using wireless sensors: (a) change in impedance signal with decreasing force level and (b) relation
between frequency shift and force loss42
ABDEL-JABER AND GLISIC 9 of 27

FIGURE 8 Change in impedance signal due to (a) changing prestress force under constant temperature and (b) changing temperature
under constant prestress force46

beam-like element with a PZT patch. The results show that damage indices from acquired impedance signals can detect
prestress loss, but quantification of losses was not investigated.43
Although impedance-based methods have shown promising short-term results, a main issue that needs to be consid-
ered before successful implementation in on-site conditions is sensitivity to varying temperature. It has been shown that
temperature can significantly impact impedance signal leading to false positive alarms in damage detection.44,45 Figure 8a
shows impedance signature under varying prestress force levels with a constant temperature, whereas Figure 8b shows
impedance signatures under a constant prestress force and varying temperatures. As shown, it is difficult to discern one
effect from the other visually. Park et al44 proposed a modified damage index to address temperature influence. This was
successfully applied in a study proposed by Huynh and Kim.46 However, it requires measurements at several different
temperatures in the healthy state in order to establish a baseline healthy impedance signal insensitive to temperature
changes. Additionally, its success largely depends on the frequency range used; Huynh et al showed that a narrower
frequency range provides better results but that could limit sensitivity to prestress loss.46
Based on the above studies, the following conclusions can be made regarding impedance-based methods for prestress
force measurement:
1. Detection of prestress losses using impedance-based methods is possible and has been achieved, as shown in
Figures 5a and 7a. Quantification of losses is only possible through calibration to develop empirical relationships
as shown in Figure 7b. Impedance signatures near the anchorage have also been shown to be insensitive to other
types of damage such as change in stiffness at locations farther away from the anchorage, which shows robustness
for detection of prestress losses.
2. Advances in sensor configuration have allowed for the reduction of the frequency range sensitive to prestress loss
from above 800 kHz (Figures 4 and 5) to under 50 kHz (Figures 6 and 7). This allows for wireless deployment of
sensors and overcomes the need for an expensive high performance impedance analyzer.
3. Temperature effects can cause changes in impedance signals that must be compensated for to accurately establish a
relationship between impedance and prestress force levels. Currently, this can only be performed through calibration
at constant force for different temperatures and using a modified damage index to account for temperature. This
allows for detection of prestress loss, but quantification is yet to be achieved.
4. The previous studies all consider prestress force change near the anchorage due to the local nature of PZT sensors.
More sensors would be required along the beam structure in order to evaluate the distribution of the prestressing
force. However, there are no examples in the literature of attachment of PZT patches to the concrete surface or
embedment along the prestressed concrete structure for the purpose of monitoring the prestress force distribution.
Thus, the possibility of monitoring the distribution of the force has not been explored.

5 ELASTO-MAGNET IC METHODS

The EM methods are based on the dependence of magnetic properties of ferromagnetic materials, such as steel, on
mechanical stress. When a ferromagnetic material is exposed to a magnetic field, a magnetic flux is induced in a pro-
cess known as magnetization. Every ferromagnetic material has a unique magnetization curve (relationship between the
10 of 27 ABDEL-JABER AND GLISIC

FIGURE 9 Typical magnetization curve52

applied magnetic field and the magnetic flux density). The ratio of the two quantities that define the magnetic curve at
any point is referred to as magnetic permeability, as given in Equation 4.47
ΔB = 𝜇ΔH, (4)
where B is the magnetic flux density, H is the applied magnetic field strength, and 𝜇 is the magnetic permeability.
A typical magnetization curve for a ferromagnetic material is given in Figure 9. As shown, the curve is nonlinear,
hysteretic, and dependent on mechanical stress.47,48 EM sensors take advantage of this nonlinearity and dependence on
stress; the value of permeability changes depending on the stress, thereby allowing for determination of stress through
measurement of magnetic permeability.49,50 To measure magnetic permeability, sensors are composed of two solenoid coils
that are installed around a steel wire or bar, thereby enclosing it without contact, that is, as a sleeve. The interrogation
unit charges the inner primary coil, which serves as the charging element and applies a magnetic field to the steel wire or
bar under consideration. The resulting magnetic flux produces an electric current measured by the outer secondary coil,
which serves as the sensing element. Relative permeability can then be calculated using the output voltage, as given in
Equation 5.51 ( )
𝜇 A0 Vout
𝜇= =1+ −1 , (5)
𝜇0 A𝑓 V0
where 𝜇 is the relative magnetic permeability of the ferromagnetic material to vacuum, 𝜇0 is the magnetic permeability
of vacuum, A0 and Af are the cross-sectional areas of the secondary coil and the steel, respectively, and Vout and V0 are the
are the induced voltage with and without the ferromagnetic material, respectively.
Because relative permeability is a property of the ferromagnetic material, steel bars or wires under consideration can
be tested in the laboratory to establish a relationship between magnetic permeability and the stress in the strand. EM
sensors show unique promise to provide measurements of absolute stress rather than relative stress even when structures
have not been instrumented during construction. The sensor was first proposed in Slovakia by Kvasnica and Fabo52 and
by Jarosevic,53 then further developed into a prototype by Sumitro et al.54 In the investigation by Kvasnica and Fabo, the
dependence of permeability on temperature was discussed and shown to be linear but also dependent on stress and the
strength of the magnetic field applied, as given in Figure 10. Sumitro et al54 performed laboratory tests using the developed
ABDEL-JABER AND GLISIC 11 of 27

FIGURE 10 Influence of temperature on magnetic permeability at different stress and magnetic field levels52

FIGURE 11 Validation of elasto-magnetic (EM) sensor measurements for a prestressed strand54

prototype on a prestressed steel bar, a single prestressed wire, and a prestressed strand and showed excellent agreement
between load cell results and EM sensor results, as shown in Figure 11.
Fabo et al first used EM sensors in several civil engineering SHM applications, including monitoring prestressing force
in unbonded tendons in a nuclear reactor envelope in Czech Republic over 5 years and monitoring the replacement
of external prestressing tendons on a bridge.55 Force determination was achieved, although comparison to other esti-
mates was not possible in the applications. The researchers concluded that an accurate determination of the magnetic
properties of the steel elements is necessary and that temperature influences on measurements can be significant. They
recommended the use of a differential sensor for temperature compensation, although details were not provided.
Wang et al further studied the relationship between magnetic permeability and stress during the prestressing of ten-
dons on a prestressed concrete bridge and showed that relative permeability varies linearly with stress.56 The researchers
confirmed the effect of temperature changes that have previously been discussed in depth in the study of Chen50 by con-
ducting a laboratory test where temperature was varied and relative permeability of steel strands was measured under
zero stress. The effects of temperature variations were shown to be linear and compensated for using empirical relation-
ships derived from the tests. Similar implementation and results were obtained by Yim et al51 through implementation
12 of 27 ABDEL-JABER AND GLISIC

on a cable-stayed bridge. A typical in situ instrumentation is shown in Figure 12. The results from the on-site calibrations
performed in the two studies are presented in Figure 13.
Another extensive calibration study was carried out by Zonta et al57 prior to the installation of EM sensors on an
in-service cable-stayed bridge. The calibration procedure was performed in two stages: laboratory calibration and on-site
calibration. The results from the laboratory calibration under constant temperature indicated that the relationship
between permeability and force is not strictly linear but better approximated by a polynomial of the third degree. How-
ever, the researchers observed that the relationship is “virtually linear” for loads above 4000 kN.57 Calibration at different
temperature confirmed the effect shown in Figure 10, where the temperature constant depends on stress. The researchers
additionally tested the repeatability of the sensor, which is particularly important for sensors assembled in situ. The
results, shown in Figure 14, indicated that although the slope of the relationship remains unchanged, the intercept varies
with different sensors. Thus, site calibration was required to ensure the correct intercept value is used. This was performed
in this study using vibration tests and the natural frequencies of the strand. However, this method yields uncertainties
that affect the accuracy of the EM sensor results.

FIGURE 12 In situ installation procedure of elasto-magnetic sensor51

FIGURE 13 On-site calibration of elasto-magnetic sensors using load cell measurements indicating (a) linear relationship between
magnetic permeability and stress for a strand that is comparable to relationships for a single wire and a strand tested in the laboratory56 and
(b) linear relationship between effective relative permeability and force51
ABDEL-JABER AND GLISIC 13 of 27

FIGURE 14 Repeatability of elasto-magnetic sensor results: voltage-force relationships acquired from three sensors assembled using the
same procedure on the same laboratory steel element showing a change in intercept57

FIGURE 15 Example of elasto-magnetic sensor installation in prestressed concrete beam prior to casting, shown with an arrow4

The previous studies focused on external prestressing tendons or strands in cable-stayed bridges. There have, how-
ever, been experiments on prestressed concrete beams, where sensors were embedded in the concrete and prestressing
forces were estimated relatively accurately for laboratory experiments.4,58 An example of such installation is presented
in Figure 15. Nevertheless, the literature lacks examples of implementation in prestressed concrete beams in field condi-
tions, possibly due to the bulky nature of the sensors and concerns regarding loss of concrete area in the vicinity of the
sensor and effects on bonding between the prestressing strands and concrete at EM sensor locations.
Based on the previous studies involving EM sensors, the following conclusions can be made:
1. EM sensors provide high sensitivity to stress via a linear relationship (as shown in Figures 13 and 14) and show
promise for absolute stress measurements, as shown in Figure 11, rather than relative stress, thereby allowing for
determination of stress levels even when sensors are installed after construction.
2. Magnetic permeability is sensitive to stress, which allows for stress measurements using EM sensors, but it is also
sensitive to temperature. As shown in Figure 10, the dependence of magnetic permeability on temperature is linear,
which simplifies analysis, but the constant varies based on the stress level. This can complicate analysis as calibration
is required at different stress levels.
3. EM sensors require calibration at known load levels. This can be performed in the laboratory on a similar steel
sample or in situ. For sensors assembled on site, in situ calibration is necessary because whereas the slope of the
linear relationship is repeatable, the intercept is not, as shown in Figure 14. In situ calibration can be challenging
in cases when the structure is already constructed and loaded.
4. Due to their bulky size (see Figures 12 and 15), EM sensors seem better suited for applications on suspension bridges
and cable-stayed bridges due to the external nature of strands. Concrete embedment has not been extensively studied
and may be challenging.
5. Although there have not been studies on the distribution of prestressing forces along a strand, magnetic permeability
appears to be a local parameter. Thus, measurement of local force changes, that is, the distribution of the force,
along a steel strand or a prestressed concrete structure would require the installation of multiple sensors along the
structure.
14 of 27 ABDEL-JABER AND GLISIC

6 ACO USTO E LAST IC M ET H OD S

Acoustoelastic methods rely on the principle of acoustoelasticity to relate the properties of ultrasonic wave propagation,
mainly wave velocity, in steel wires and strands to applied stress.59 The dependence of the change in wave velocity on
stress changes has been utilized in several applications in civil engineering including measurement of residual stress60,61
and of bolt stress.62,63
The first attempt to study elastic wave propagation in prestressed steel strands was an experimental study conducted by
Kwun et al in 1998, where the researchers tested specimens of prestressing strands instrumented with magnetostrictive
sensors, to emit and detect elastic waves, under different stress levels.64 The main finding from the testing was that the
presence of tensile load on the specimens caused significant attenuation in a specific portion of the frequency components
of the wave, which resulted in the disappearance of this portion. The center frequency of the portion (notch frequency)
was found to be linearly dependent on the logarithm of the applied load, as shown in Figure 16. Despite the absence of
a theoretical basis for this result, it indicated promise in the use of longitudinal wave properties for the determination of
prestress forces.
Several future attempts focused on the time of flight of waves (or the wave velocity) between an input and output trans-
ducer to quantify changes in stress, based on theoretical analysis of the wave velocity. In a homogeneous isotropic material
under uniform axial stress, the bulk longitudinal plane wave velocity can be written in the first-order approximation as
given in Equation 6.59,62,65,66
√ [ ( )]
𝜆 + 2𝜇 𝜎 𝜆+𝜇
V𝜎 = 1+ (4𝜆 + 10𝜇 + 4m) + 𝜆 + 2l , (6)
𝜌 2(𝜆 + 2𝜇)(3𝜆 + 2𝜇) 𝜇

where 𝜌 is the mass density, 𝜎 is the tensile stress, 𝜆 and 𝜇 are Lame's elastic constants, and m and l are Murnaghan's third
order elastic constants. Equation 6 can be rewritten as follows:

V𝜎 = V0 (1 + K𝜎), (7)

𝜆+2𝜇
where V0 = 𝜌
is the longitudinal wave velocity under zero stress, and K = 1
( 𝜆+𝜇 (4𝜆 + 10𝜇 + 4m) + 𝜆 + 2l).
2(𝜆+2𝜇)(3𝜆+2𝜇) 𝜇

FIGURE 16 Linear relationship between notch frequency and applied tensile load64
ABDEL-JABER AND GLISIC 15 of 27

A study was conducted by Chen and Wissawapaisal to estimate the wave velocity of a 4.76-m long prestressing strand.67,68
The instrumentation employed in the studies consisted of one AE transducer at each end of the prestressing strand, as
shown in Figure 17. Sample results are presented in Figure 18 and indicate a linear increase in the time of flight for stress
levels between 18% and 70% of the ultimate tensile strength (UTS) of the strand. Outside of this range, the dependence
of wave velocity on stress levels tends to be nonlinear. Figure 18 also shows good agreement between experimental and
analytical results obtained using Equation 6. Another important finding from the two studies is the detection of waveforms
for longer steel strands (up to 83.94 m), indicating the possibility of applying the method to full-scale prestressed concrete
structures with ungrouted tendons.68
Other researchers recognized the difficulty in estimating the parameters required to determine the acoustoelastic factor,
K, but were interested in experimentally quantifying the parameter as an indication of the sensitivity of the wave velocity
for stress measurements. If incremental stress changes are considered instead of absolute stress, the incremental velocity
change as a percentage of the velocity at a known reference stress, V𝜎R , is given by Equation 8, and K denotes the sensitivity
of the relation.69
ΔV𝜎
= KΔ𝜎. (8)
V𝜎R

FIGURE 17 Instrumentation of an ungrouted tendons in prestressed concrete beam with AE transducers, (a) schematic illustration,67 and
(b) laboratory deployment68

FIGURE 18 Time shift comparison between experimental (dotted lines) and analytical results (solid lines) for the fourth cycle waveform67
16 of 27 ABDEL-JABER AND GLISIC

Washer et al69 performed tests on strands and center wires from decomposed strands to experimentally determine the
acoustoelastic factor. The tests indicated that the relationship is not linear throughout the full range of tensile stresses.
It is, however, linear for stress ranges with lower bounds between 30% and 50% of UTS and upper bounds between 70%
and 80%, which is acceptable for typical stress level in strands in prestressed concrete structures. The constant K in this
range was determined to be less than 1% per GPa of stress for all tested specimens, as shown in Table 1. Additionally,
the tests indicated that the constant is different even for strands that meet the same specifications but manufactured at
different times, as shown in Figure 19, indicating the need for calibration on the same strands used in a project prior
to implementation on site. The low sensitivity observed in the study raises doubts regarding monitoring relatively low
prestress losses throughout the service life of a structure. One subsequent study showed that the change in the wave
velocity as a function of applied stress was nonlinear in stress levels below 20% of UTS of the strand, but linear throughout
the entire range for a single core wire,70 whereas another showed nonlinearity at low stress levels even in the single wire.66
The acoustoelastic coefficient was different in the studies but still under 1% per GPa. Additionally, the previous three
studies all discuss the potential of increasing the sensitivity of the method by including the effect of the elongation of the
strand under stress, that is, by clamping the sensors to the strand and allowing the distance between them to increase
with increasing stress. Although the elongation effect increases sensitivity, as shown in Figure 20, it remains under 1%
per GPa.66,69,70
Besides wave velocity, researchers have studied other properties of wave propagation, such as the transmitted ultrasonic
energy in terms of the amplitude of the frequency peak and the area under the energy spectrum.71,72 The relationships of
these properties with load are shown in Figure 21 for a laboratory test performed by Rizzo on a core wire.72 Both parameters
were found to linearly decrease with increasing tensile stress for a single wire, with a significant decrease in the area under
the energy spectrum of 45% per GPa.72 When applied to an assembled strand, the results did not show linear behavior,
leading the author to recommend instrumentation of the core wire only. It should be noted that the sensing mechanism
requires direct access to the ends of the strands, thereby limiting the application to shorter strands due to attenuation
concerns, up to a few meters long.72 Further investigations or site implementations of this method have not been reported.
All the aforementioned studies have been on ungrouted strands or wires due to the effect of grouting material on atten-
uation, thereby restricting the applications for prestressed concrete to unbonded tendons. Additionally, studies have not
investigated the effect of temperature on measurements, as well as the possibility of embedding sensors in the concrete.

TABLE 1 Base velocities measured at 50% of the ultimate tensile strength and velocity
constants (acoustoelastic factor) for wires C1–C4 and rod R1, reproduced from the study of 69
Material Stress range Phase velocity Group velocity kp kg
designation (GPa) (m/s) (m/s) (%/GPa) (%/GPa)
C1 0.90–1.58 5,061 4,892 -0.68 -0.69
C2 5,119 4,941 -0.50 -0.52
C3 0.90–1.58 5,060 4,715 -0.59 -0.62
C4 5,006 4,668 -0.62 -0.65
R1 0.5–1.23 4,937 4,596 -0.74 -0.69

FIGURE 19 Change in velocity with change in stress for two center wires manufactured at different times with the same specifications69
ABDEL-JABER AND GLISIC 17 of 27

FIGURE 20 Effect of elongation on the sensitivity of time delay measurements on applied stress70

FIGURE 21 Load dependence on (a) transmitted energy peak amplitude and (b) transmitted energy area72

Based on the previous studies, the following conclusions can be made regarding acoustoelastic methods for measuring
prestress forces:
1. Velocity of stress waves in prestressing strands is linearly dependent on stress for typical stress levels in prestressing
strands, enabling measurement using the acoustoelastic effect. However, change in velocity shows inherently low
sensitivity to stress changes, less than 1% per GPa, as shown in Table 1. While including the effect of strand elongation
can improve sensitivity, as shown in Figure 20, it remains low compared with other methods.
18 of 27 ABDEL-JABER AND GLISIC

2. Absolute measurement of stress is possible as shown in Figure 18 but requires approximating the parameters in
Equation 6, which can be challenging. Instead, relative or absolute stress can be determined through laboratory
calibration and Equation 8, as shown in Table 1 and Figure 19.
3. Issues such as the effect of temperature on measurements, as well as attenuation due to the presence of grouting
material, challenge the application of the methods in field settings.
4. The nature of wave propagation methods is global; wave velocities reflect the global behavior of a strand, rather than
behavior in the vicinity of the transducer. Therefore, monitoring the distribution of the prestressing force along a
prestressed concrete structure is currently not achievable using acoustoelastic methods.

7 STRAIN-BASED METHODS

Due to the well-established relationship between strain and stress in the linear elastic range, strain monitoring is regarded
as the closest to direct monitoring of stress. Recent advances in the development of durable and stable strain sensors have
led to the increasing popularity and use of strain-based methods for prestress force monitoring. In fact, the ACI 423 Guide
to Estimating Prestress Loss7 lists measured prestress loss values in laboratory or field setting from more than 40 studies,
all of which are conducted using types of strain sensors.
If structures are instrumented with sensors during construction, either through attaching them to prestressing strands
or embedding them in the concrete, long-term strain changes can be monitored. Additionally, instrumentation during
construction allows for measuring strain changes during prestress force transfer, thereby enabling calculation of the initial
stress in the prestressing strands. If the strain change at the location of the prestressing strands is determined, the stress
change can be calculated using Hooke's law, given by Equation 9.

Δ𝜎p = EΔ𝜀p , (9)

where Δ𝜎 p is the average stress change in the prestressing strand, E is the modulus of elasticity of the prestressing steel,
and Δ𝜀p is the strain change at the location of the center of gravity of the prestressing strands.
In cases where the location of the prestressing strand is not known with accuracy or where there is concerns regarding
separating the effects of bending due operational load on the structure from the effects of prestress loss, strain changes at
the centroid of stiffness of the concrete cross-section can be monitored instead of strain changes at the center of gravity
of prestressing strands.73 Because strain at the centroid of stiffness is minimally affected by bending due to load and
temperature changes, the effect of long-term prestress losses can be isolated.
In applications where the strain sensors are embedded in the concrete, at least two strain sensors are commonly embed-
ded in every instrumented cross section. Assuming a linear strain distribution, the strain at the center of gravity of the
prestressing strand can be determined. Assuming a perfect bond between the prestressing strands and surrounding con-
crete, the strain changes measured in the concrete can be assumed to be equal to the strain changes in the prestressing
strands at the same location, and Equation 9 can be used to estimate the stress change, as illustrated in Figure 22. This
type of instrumentation has been used by many researchers employing vibrating wire strain gauges5,74-77 (see Figure 23a),
demountable mechanical gauges,78,79 electrical strain gauges,80 or fiber optic sensors (see Figure 23b).2,83
Alternatively, strain sensors can be placed at the location of the centroid of the prestressing strands84 or attached directly
to prestressing strands to reduce the number of required sensors and directly measure the strain changes at or in the
steel strand. Such applications have utilized similar types of sensors such as electrical resistance strain gauges85,86 (see
Figure 23c), distributed fiber optic sensors (see Figure 23d),87 and discrete fiber optic sensors.88 However, attaching strain
sensors directly to strands exposes them to large tensile strains during prestressing, which can cause damage to the sen-
sors. Some researchers attempted to overcome the issue by embedding the sensors into metal packaging prior to attaching
them.82 This approach, shown in Figure 23d, has only been tested in the laboratory and some sensors were damaged in
the process regardless of the packaging, indicating that the packaging needs to be improved.
Two examples of results are presented in Figure 24. Figure 24a shows results from a 60-day laboratory experiment where
prestress loss in the tendons is measured using strain sensors and compared with design code estimates. Figure 24b shows
results from a field instrumentation on a posttensioned pedestrian bridge, where prestress loss is measured over a period
of seven years, confirming the longevity of sensors in field conditions. Additionally, using a set of discrete sensors along
the length of a structure or a distributed sensor, the distribution of the prestressing force along the structure at a given
point in time can be determined, as shown in Figure 25.
ABDEL-JABER AND GLISIC 19 of 27

FIGURE 22 Typical strain sensor instrumentation and strain distribution76

FIGURE 23 Different types of strain sensors used in monitoring concrete: (a) vibrating wire gauge,76 (b) distributed and discrete fiber
optic sensors,73 (c) electrical strain gauge,81 and (d) metal-coated fiber optic sensor82

Although the instrumentation approaches discussed above have been applied extensively in field settings, newer
approaches are still in the laboratory-testing phase. One such approach is replacing the core wire of a seven-wire pre-
stressing strand by a fiber-reinforced polymer where an optical fiber can be embedded during manufacturing for sensing
purposes, as shown in Figure 26.89,90 Although laboratory tests indicate that the embedded sensor can provide accurate
strain measurements, the application in field settings is yet to be explored.
20 of 27 ABDEL-JABER AND GLISIC

FIGURE 24 Example of prestress loss measured using strain sensors from (a) laboratory experiment lasting 60 days86 and (b) field
measurements collected over seven years73

FIGURE 25 Distribution of prestressing force along a structure using (a) distributed fiber optic sensor87 and (b) set of discrete fiber optic
sensors

FIGURE 26 Smart prestressing strand with a carbon core wire and embedded fiber optic sensor89
ABDEL-JABER AND GLISIC 21 of 27

A common challenge in strain-based monitoring is the sensitivity of strain measurements to temperature changes in
terms of the thermal effects on both the sensor and the structure. However, several commercial sensors currently exist
on the market, which have been tested for temperature influences and manufacturers provide calibrated constants for
temperature compensation. As for thermal effects on the structure, they can be sufficiently accurately compensated for
using Equation 10.
𝜀compensated = 𝜀measured − 𝛼T ΔT, (10)
where 𝜀compensated is the thermally compensated strain, 𝜀measured is the measured strain, 𝛼 T is the thermal expansion coef-
ficient of the steel or the concrete (depending on the material the sensor is attached to), and ΔT is the temperature
change.
Thus, temperature monitoring is required in field conditions. The thermal expansion coefficient can be determined in
the field or during laboratory testing.
Strain-based methods show promise for monitoring prestressing losses as demonstrated above if structures are instru-
mented during construction because estimation of stress changes is based on changes in strain and not absolute
measurement of stress. However, if structures are not instrumented during construction, semidestructive or destructive
strain-based methods can be used to determine the remaining prestress force as demonstrated by some studies. Tabatabai
and Dickson91 performed extensive testing on a 34-year old posttensioned concrete girder acquired after dismantling of
a bridge. First, they loaded the beam until the first crack occurred then unloaded it and instrumented the crack with
a crack gauge. By reloading the beam until the crack width began to increase, they were able to determine the bend-
ing moment at which decompression occurred. This allowed for back-calculating the prestress force. This technique has
been used on several older beams that were tested to failure.92-94 A less invasive method was used by other researchers,
which involves instrumenting wires from prestressing strands with strain sensors and cutting the wires to observe the
strain release, shown in Figure 27, which can be used to infer the stress in the wire prior to cutting, using Equation 9.81,95
Remennikov and Kaewunruen reported that less than 5% of the concrete along a beam structure needs to be removed to
expose the prestressing strands and allow for such testing.81 Additionally, due to the friction between wires, this will only
have a minor effect on the prestressing force locally.
Based on the above studies utilizing strain measurements, the following conclusions can be made regarding strain-based
assessment of prestressing forces:
1. Strain monitoring is a mature sensing field with a well-defined relationship with stress, given by Equation 9, which
allows for monitoring prestressing forces in field applications. However, instrumentation during construction is
required because measured changes are relative and not absolute.
2. Several applications of strain-based monitoring of prestressing forces have been reported utilizing different types of
strain sensors, as shown in Figure 23. Newer technologies involve smart strands that provide the required strength,
as well as sensing capabilities, as shown in Figure 26, are currently in the laboratory development phase.

FIGURE 27 Semidestructive strain-based determination of prestressing force: (a) cutting of wires in prestressed concrete beam and (b)
strain change during cutting process81
22 of 27 ABDEL-JABER AND GLISIC

3. Sufficiently accurate temperature compensation for strain measurements is well developed, as given by Equation 10
and has allowed for field monitoring, even under varying temperatures.
4. Because strain is a local parameter, it measures effects in the vicinity of the sensor, thereby allowing for the mea-
surement of the distribution of the prestressing force along a structure using multiple discrete sensors (Figure 25b),
or a distributed sensor (Figure 25a).
5. Destructive or semidestructive techniques can be used to determine prestressing forces in structures that have not
been instrumented during construction, as shown in Figure 27. This is more applicable in laboratory settings for
research purposes or in rehabilitation projects.
In addition to the numerous applications of strain-based monitoring of prestressing forces, a formalized method for
design of the sensor network, analysis of collected data, and assessment of uncertainties was proposed in the study of
Abdel-Jaber and Glisic96 and applied to a structure instrumented during construction. The resulting prestressing force
distribution and associated uncertainties are presented in Figure 25b.

8 CO NCLUSIONS AND RECO MMENDATIONS

Five different classes of methods for monitoring prestress losses were presented and discussed in the previous five sections.
Each section ends with concise conclusions about corresponding techniques written in form of bullet points. These con-
clusions are summarized and compared in this section, and complemented with identified potential directions for future
research. The following conclusions, also summarized in Table 2, can be made regarding the promise of each class of
methods for monitoring prestressed concrete structures:
1. Vibration methods do not show promise for monitoring prestressed concrete structures due to the contradicting
arguments regarding the effect of the prestressing force change on natural frequencies, particularly in uncracked
beams. Additionally, the global nature of vibration monitoring further limits their applicability; other factors such as
changes in boundary conditions, environmental conditions, and other types of damage can cause significant changes
in natural frequencies, making it difficult to discern or separate the effects of prestress losses, if they are detected.
The authors could not immediately identify future research in vibration methods that could be helpful for reliable
detection and quantification of prestress losses.
2. Impedance-based methods show promise for detecting local changes in prestressing forces. They do, however,
require calibration and are sensitive to temperature changes, which might necessitate further calibration under
varying temperatures. Due to the local nature of the sensors and methods, current applications only provide infor-
mation about the prestressing force near the anchorage, where the sensor is typically installed. Future research in
impedance-based methods could focus on methods for reliable quantification of prestress losses, methods for imple-
mentation on concrete, if possible, (surface or embedded), and enabling monitoring of force distribution along the
structure, that is, at multiple points along the structure (as opposed to a limited number of points at anchorage
locations).
3. EM methods tend to be more suitable for external tendons due to the bulkiness of sensors, which complicates embed-
ment in concrete structures where the density of reinforcement can affect sensor placement. They do, however, show

TABLE 2 Comparison of methods for monitoring of prestressing forces


Vibration Impedance Elasto-magnetic Acoustoelastic Strain
Sensitivity to prestress Debatable High (locally) High Low High
force changes
Effects of environmental High High High (can be Not explored High (can be
factors calibrated) compensated)
Distribution of force Not feasible, global Local but only Feasible, local Not feasible, Feasible, local
along the structure at anchorage global
Feasibility of instrumentation High High Low High High
Applicability to real-life Unfeasible Feasible with Feasible with Feasible with Feasible
structures calibration calibration attenuation
considerations
ABDEL-JABER AND GLISIC 23 of 27

promise for assessment of absolute force levels in external tendons, even when structures are not instrumented dur-
ing construction. In theory, although calibration is required, it can be performed on a sample of prestressing strand
in the laboratory and does not need to be performed on site. However, repeatability of sensors built in situ requires
knowledge of the initial stress in the strand for calibration. Future research in EM methods could focus on pack-
aging and methodology for embedding the sensors in concrete, which would enable monitoring distribution of the
force along the structure.
4. Acoustoelastic methods show low sensitivity to prestressing force changes. The issue is compounded by attenuation
of stress waves, which limits their application to ungrouted, shorter tendons because access to both ends of a tendon
is required. Future research could focus on increasing the sensitivity of the method in grouted tendons, evaluating
and compensating temperature effects, and creating an embeddable sensor, which would enable monitoring the
distribution of the force along the structure.
5. Strain-based methods show the most promise with regard to monitoring prestressing forces due to the maturity of
the sensing technologies and the well-established relationship between strain and stress. The only limitation is the
requirement for instrumentation at construction because strain changes are relative and not absolute. Given that
the strain-based methods are relatively mature both in terms of sensors and algorithms, future research could focus
on pervasive implementation of strain sensors, for example, by using distributed fiber optic sensors, self-sensitive
(nano-) materials, large-area electronics, and so on, and algorithms that combine mathematical/numerical models
with monitoring results, to more accurately discern prestress losses from other unusual behaviors.
Despite the existence of a wide variety of methods for monitoring of prestressing forces in prestressed concrete struc-
tures, only strain-based methods are extensively used in field applications. Additionally, with the exception of a few
studies,96 the literature lacks guidelines for the design of sensor networks or assessment of uncertainties. Such guidelines
are required to establish comparisons between results from different studies, as well as to effectively assess results from
the monitoring systems.
Applications of the other classes of methods need to be extended to field conditions in conjunction with reliable
strain-based methods to validate such sensors and methods. As demonstrated previously, impedance-based and EM meth-
ods show promise to overcome the shortcomings of the widely used strain-based methods, such as the requirement for
installation during construction. Thus, their extension to field applications can provide assessment for existing prestressed
concrete infrastructure elements. Aspects of their application, such as longevity in field settings, and accurate calibration
are yet to be investigated in detail with respect to prestressed concrete structures, as opposed to strands in suspension and
cable-stayed bridges.
Other directions for future research include implementing current strain-based methods, such as smart strands dis-
cussed in strain-based methods, in field settings and comparing their performance to steel tendons. Research should
investigate the cost of using such technologies in the field, their robustness to harsh site conditions, and their long-term
performance.

ACKNOWLEDGEMENTS
This work was partially supported by the National Science Foundation grant no. CMMI-1362723. Any opinions, find-
ings, and conclusions or recommendations expressed in this paper are those of the authors and do not necessarily reflect
the views of the National Science Foundation. The work was also supported by USDOT Office of the Assistant Secre-
tary for Research and Technology grant no. DTRT13-G-UTC28. The views, opinions, findings, and conclusions are the
responsibility of the authors only and do not represent the official policy or position of the USDOT or any state or entity.

ORCID

Hiba Abdel-Jaber https://orcid.org/0000-0001-7044-3622

REFERENCES
1. Federal Highway Administration. NBI ASCII Files. https://www.fhwa.dot.gov/bridge/nbi/ascii.cfm?year=2016; 2016.
2. Idriss R, Solano A. Effects of steam curing temperature on early prestress losses in high-performance concrete beams.
Transp Res Rec: J Transp Res Board. 2002;1813(1):218-228. http://trb.metapress.com/index/d39134tx838m34x8.pdf%5Cn
http://trb.metapress.com/openurl.asp?genre=article&id=doi:10.3141/1813-26 http://trrjournalonline.trb.org/doi/10.3141/1813-26
24 of 27 ABDEL-JABER AND GLISIC

3. Kahn LF, Lopez M. Prestress losses in high performance lightweight bridge girders. PCI J. 2005;50(5):84-94.
4. Halvonik J, Dolnak J, Borzovic V. Long-term losses of prestress in precast members cast from HPC. Procedia Eng. 2013;65:81-86.
5. Stallings IM, Barnes RW, Eskildsen S. Camber and prestress losses in Alabama HPC bridge girders. PCI J. 2003;48(5):90-104.
http://www.pci.org/pci_journal-2003-september-october-9/
6. He W. Creep and shrinkage of high performance concrete and prediction of the long-term camber of prestressed bridge girders. Ph.D.
Thesis: Iowa State University; 2013.
7. Joint ACI-ASCE Committee 423. Guide to estimating prestress loss, Farmington Hills, MI, American Concrete Institute; 2016.
8. Woodward RJ, Williams FW. Collapse of Ynys-y-Gwas bridge, West Glamorgan. Proc Inst Civ Eng Civ Eng. 1988;84(4):635-669.
9. Watson SC, Stafford D. Cables in trouble. Civ Eng. 1988;58(4):38-41. http://www.scopus.com/inward/record.url?eid=2-s2.0-0023817032&
partnerID=tZOtx3y1
10. Abraham MA, Park S, Stubbs N. Loss of prestress prediction based on nondestructive damage location algorithms. Smart Structures and
Materials 1995: Smart Systems for Bridges, Structures, and Highways, Vol. 2446. San Diego, CA: SPIE: The international society for optics
and photonics; 1995:60-67.
11. Casas JR. A combined method for measuring cable forces: The cable-stayed alamillo bridge, spain. Struct Eng Int. 1994;4(4):235-240.
https://doi.org/10.2749/101686694780601700
12. Zui H, Shinke T, Namita Y. Practical formulas for estimation of cable tension by vibration method. J Struct Eng. 1996;122(6):651-656.
13. Yang Y, Li S, Nagarajaiah S, Li H, Zhou P. Real-time output-only identification of time-varying cable tension from accelerations via
complexity pursuit. J Struct Eng. 2016;142(1):04015083.
14. Saiidi M, Douglas B, Feng S. Prestress force effect on vibration frequency of concrete bridges. J Struct Eng. 1994;120(7):2233-2241.
15. Miyamoto A, Tei K, Nakamura H, Bull JW. Behavior of prestressed beam strengthened with external tendons. J Struct Eng.
2000;126(9):1033-1044. http://ascelibrary.org/doi/10.1061/%28ASCE%290733-9445%282000%29126%3A9%281033%29
16. Kim JT, Yun CB, Ryu YS, Cho HM. Identification of prestress-loss in PSC beams using modal information. Struct Eng Mech.
2004;17:467-482. http://multi-science.atypon.com/doi/10.1260/1369-4332.15.6.997http://koreascience.or.kr/journal/view.jsp?kj=KJKHB9
&py=2004&vnc=v17n3_4&sp=467
17. Noble D, Nogal M, O'Connor A, Pakrashi V. The effect of post-tensioning force magnitude and eccentricity on the natural bending fre-
quency of cracked post-tensioned concrete beams. J Phys: Conf Ser. 2015;628:012047. http://stacks.iop.org/1742-6596/628/i=1/a=012047?
key=crossref.ff6ecf811ce98dedf02e04d9f2466744
18. Grace NF, Ross B. Dynamic characteristics of post-tensioned girders with web openings. J Struct Eng. 1996;122(6):643-650.
19. 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(3):210-220.
20. Hamed E, Frostig Y. Free vibrations of cracked prestressed concrete beams. Eng Struct. 2004;26(11):1611-1621.
21. De Roeck G. The state-of-the-art of damage detection by vibration monitoring: The SIMCES experience. J Struct Control.
2003;10(2):127-134.
22. DaIl'Asta A, Dezi L. Discussion on prestress force effect on vibration frequency of concrete bridges. J Struct Eng. 1996;122(4):458.
23. Deak G. Discussion on prestress force effect on vibration frequency of concrete bridges. J Struct Eng. 1996;122(4):458-459.
24. Jain SK, Giel SC. Discussion on prestress force effect on vibration frequency of concrete bridges. J Struct Eng. 1996;122(4):459-460.
25. 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. https://doi.org/10.1016/j.jsv.2015.06.021
26. Hamed E, Frostig Y. Natural frequencies of bonded and unbonded prestressed beams–prestress force effects. J Sound Vib.
2006;295(1-2):28-39. http://linkinghub.elsevier.com/retrieve/pii/S0022460X06000915
27. 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.
28. Huynh TC, Park YH, Park JH, Hong DS, Kim JT. Effect of temperature variation on vibration monitoring of prestressed concrete girders.
Shock Vib. 2015;2015:1159-1175.
29. Peeters B, De Roeck G. One-year monitoring of the Z24-Bridge: Environmental effects versus damage events. Earthquake Eng Struct Dyn.
2001;30(2):149-171. http://doi.wiley.com/10.1002/1096-9845%28200102%2930%3A2%3C149%3A%3AAID-EQE1%3E3.0.CO%3B2-Z
30. Cornwell P, Farrar CR, Doebling SW, Sohn H. Environmental variability of modal properties. Exp Tech. 1999;23(6):45-48. http://doi.wiley.
com/10.1111/j.1747-1567.1999.tb01320.x
31. Park G, Sohn H, Farrar CR, Inman DJ. Overview of piezoelectric impedance-based health monitoring and path forward. The Shock Vib
Digest. 2003;35(6):451-463. http://svd.sagepub.com/cgi/doi/10.1177/05831024030356001
32. Liang C, Sun FP, Rogers CA. Coupled electro-mechanical analysis of adaptive material systems—Determination of the actuator
power consumption and system energy transfer. J Intell Mater Syst Struct. 1994;5(1):12-20. http://jim.sagepub.com/cgi/doi/10.1177/
1045389X9400500102
33. Kim JT, Park JH, Hong DS, Cho HM, Na WB, Yi JH. Vibration and impedance monitoring for prestress-loss prediction in PSC girder
bridges. Smart Struct Syst. 2009;5(1):81-94. http://koreascience.or.kr/journal/view.jsp?kj=KJKHFZ&py=2009&vnc=v5n1&sp=81
34. Kim JT, Park JH, Hong DS, Park WS. Hybrid health monitoring of prestressed concrete girder bridges by sequential vibration-impedance
approaches. Eng Struct. 2010;32(1):115-128. https://doi.org/10.1016/j.engstruct.2009.08.021
35. Soh CK, Tseng K, Bhalla S, Gupta A. Performance of smart piezoceramic patches in health monitoring of a RC bridge. Smart Mater Struct.
2000;9:533-542.
ABDEL-JABER AND GLISIC 25 of 27

36. Raju V, Park G, Cudney H. Impedance-based health monitoring technique of composite reinforced structures. In: Proceedings of the 9th
International Conference on Adaptive Structures and Technologies; 1998; Cambridge, MA:448-457.
37. Chaudhry Z, Lalande F, Ganino A, Rogers C, Chung J. Monitoring the integrity of composite patch structural repair via piezoelectric
actuators/sensors. 36th structures, structural dynamics and materials conference. Reston, Virigina: American Institute of Aeronautics and
Astronautics; 1995. http://arc.aiaa.org/doi/10.2514/6.1995-1074
38. Bois C, Hochard C. Monitoring of laminated composites delamination based on electro-mechanical impedance measurement. J Intell
Mater Syst Struct. 2004;15(1):59-67. http://journals.sagepub.com/doi/10.1177/1045389X04039405
39. Xu YG, Liu GR. A modified electro-mechanical impedance model of piezoelectric actuator-sensors for debonding detection of composite
patches. J Intell Mater Syst Struct. 2002;13(6):389-396. http://journals.sagepub.com/doi/10.1177/104538902761696733
40. Abe M, Park G, Inman DJ. Impedance-based monitoring of stress in thin structural members. In: Proceedings of the 11th International
Conference on Adaptive Structures and Technologies; 2000; Nagoya, Japan:285-292.
41. Park JH, Lee SY, Hong DS, Kim JT. Prestress-force monitoring of PSC girder bridges using wireless impedance sensor nodes. Proceedings
of SPIE—Sensors and Smart Structures Technologies for Civil, Mechanical, and Aerospace Systems, Vol. 7647. San Diego, CA: SPIE: The
international society for optics and photonics; 2010. http://proceedings.spiedigitallibrary.org/proceeding.aspx?articleid=757859
42. Nguyen KD, Kim JT. Smart PZT-interface for wireless impedance-based prestress-loss monitoring in tendon-anchorage connection. Smart
Struct Syst. 2012;9(6):489-504. http://ebooks.cambridge.org/ref/id/CBO9781107415324A009
43. Huynh TC, Kim JT. Impedance-based cable force monitoring in tendon-anchorage using portable PZT-interface technique. Math Prob
Eng. 2014;2014:1-11. http://www.hindawi.com/journals/mpe/2014/784731/
44. Park G, Kabeya K, Cudney HH, Inman DJ. Impedance-based structural health monitoring for temperature varying applications. JSME Int
J Ser A. 1999;42(2):249-258. http://joi.jlc.jst.go.jp/JST.Journalarchive/jsmea1997/42.249?from=CrossRef
45. Koo KY, Lee JJ, Yun CB, Kim JT. Damage detection in beam-like structures using deflections obtained by modal flexibility matrices. Adv
Sci Technol. 2008;56:483-488. http://www.scientific.net/AST.56.483
46. Huynh T-C, Kim J-T. Quantification of temperature effect on impedance monitoring via pzt interface for prestressed tendon anchorage.
Smart Mater Struct. 2017;26(12):125004. http://stacks.iop.org/0964-1726/26/i=12/a=125004
47. Bertotti G. Hysteresis in magnetism: For physicists, materials scientists, and engineers: Academic Press; 1998. http://linkinghub.elsevier.
com/retrieve/pii/B9780120932702500714
48. Cullity BD, Graham CD. Introduction to magnetic materials. 2nd ed. New Jersey: Wiley; 2008.
49. Wang G, Wang ML. The utilities of U-shape EM sensor in stress monitoring. Struct Eng Mech. 2004;17(3_4):291-302.
http://koreascience.or.kr/journal/view.jsp?kj=KJKHB9&py=2004&vnc=v17n3_4&sp=291
50. Chen ZL. Characterization and constitutive modeling of ferromagnetic materials for measurement of stress. Ph.D. Thesis: The University
of Illinois at Chicago; 2000.
51. Yim J, Wang ML, Shin SW, et al. Field application of elasto-magnetic stress sensors for monitoring of cable tension force in cable-stayed
bridges. Smart Struct Syst. 2013;12:465-482.
52. Kvasnica B, Fabo P. Highly precise non-contact instrumentation for magnetic measurement of mechanical stress in low-carbon steel wires.
Meas Sci Technol. 1996;7:763-767.
53. Jarosevic A. Magnetoelastic method of stress measurement in steel. Dordrecht: Springer Netherlands; 1999;107-114.
https://doi.org/10.1007/978-94-011-4611-1_13
54. Sumitro S, Jarosevic A, Wang ML. Elasto-magnetic sensor utilization on steel cable stress measurement. In: The First Fib Congress,
Concrete Structures in the 21th Century; 2002; Osaka:13-19.
55. Fabo P, Jarosevic A, Chandoga M. Health monitoring of the steel cables using the elasto-magnetic method. Proceedings of IMECE2002
ASME International Mechanical Engineering Congress & Exposition. New Orleans, Louisiana: ASME; 2002:295-299.
56. Wang ML, Wang G, Zhao Y. Application of EM stress sensors in large steel cables. Sensing Issues in Civil Structural Health Monitoring.
Berlin/Heidelberg: Springer-Verlag; 2005:145-154. http://link.springer.com/10.1007/1-4020-3661-2_15
57. Zonta D, Esposito P, Molignoni M, et al. Calibration of elasto-magnetic sensors for bridge-stay cable monitoring. Proc 6th Eur Workshop
Struct Health Monit. 2012:1-8.
58. Sumitro S, Kurokawa S, Shimano K, Wang ML. Monitoring based maintenance utilizing actual stress sensory technology. Smart Mater
Struct. 2005;14(3):S68-S78. http://stacks.iop.org/0964-1726/14/i=3/a=009?key=crossref.da7a13a676dcc3d6924dd04f3eb16a72
59. Hughes DS, Kelly JL. Second-order elastic deformation of solids. Phys Rev. 1953;92(5):1145-1149. http://link.aps.org/doi/10.1103/PhysRev.
92.1145
60. Blinka J, Sachse W. Application of ultrasonic-pulse-spectroscopy measurements to experimental stress analysis. Exp Mech.
1976;16(12):448-453. http://link.springer.com/10.1007/BF02324101
61. Allen DR, Sayers CM. The measurement of residual stress in textured steel using an ultrasonic velocity combinations technique. Ultrason.
1984;22(4):179-188. http://linkinghub.elsevier.com/retrieve/pii/0041624X84900349
62. Johnson GC, Holt AC, Cunningham B. An ultrasonic method for determining axial stress in bolts. J Test Eval. 1986;14(5):253. http://www.
astm.org/doiLink.cgi?JTE10337J
63. Bickford JH. Using ultrasonics to measure the residual tension in bolts. Exp Tech. 1988;12(11):3s-5s.
64. Kwun H, Bartels KA, Hanley JJ. Effects of tensile loading on the properties of elastic-wave propagation in a strand. The J Acoust Soc Am.
1998;103(6):3370.
65. Murnaghan FD. Finite Deformation of an Elastic Solid. New York: Wiley; 1951. http://www.sciencemag.org/cgi/doi/10.1126/science.115.
2997.634-a
26 of 27 ABDEL-JABER AND GLISIC

66. Chaki S, Bourse G. Stress level measurement in prestressed steel strands using acoustoelastic effect. Exp Mech. 2009;49(5):673-681. http://
link.springer.com/10.1007/s11340-008-9174-9
67. Chen HLR, Wissawapaisal K. Measurement of tensile forces in a seven-wire prestressing strand using stress waves. J Eng Mech.
2001;127(6):599-606. http://ascelibrary.org/doi/10.1061/%28ASCE%290733-9399%282001%29127%3A6%28599%29
68. Chen HLR, Wissawapaisal K. Application of Wigner-Ville transform to evaluate tensile forces in seven-wire prestressing strands. J Eng
Mech. 2002;128(11):1206-1214. http://ascelibrary.org/doi/10.1061/{%}28ASCE{%}290733-9399{%}282002{%}29128{%}3A11{%}281206{%}29
69. Washer GA, Green RE, Pond RB. Velocity constants for ultrasonic stress measurement in prestressing tendons. Res Nondestr Eval.
2002;14(2):81-94. http://www.tandfonline.com/doi/abs/10.1080/09349840209409706
70. Lanza di Scalea F, Rizzo P, Seible F. Stress measurement and defect detection in steel strands by guided stress waves, Vol. 15; 2003:219-227.
http://ascelibrary.org/doi/10.1061/%28ASCE%290899-1561%282003%2915%3A3%28219%29
71. Rizzo P, Lanza di Scalea F. Wave propagation in multi-wire strands by wavelet-based laser ultrasound. Exp Mech. 2004;44(4):407-415.
http://exm.sagepub.com/cgi/doi/10.1177/0014485104044321
72. Rizzo P. Ultrasonic wave propagation in progressively loaded multi-wire strands. Exp Mech. 2006;46(3):297-306.
73. Abdel-Jaber H. Comprehensive strain-based methods for monitoring prestressed concrete beam-like elements. Ph.D. Thesis: Princeton
University; 2017.
74. Barr P, Eberhard M, Stanton J, Khaleghi B, Hsieh JC. High performance concrete in Washington state SR18/SR516 overcrossing: Final
report on girder monitoring, Washington State Transportation Center; 2000.
75. Kamatchi P, Dhayalini B, Balaji Rao K, Saibabu S, Parivallal S, Ravisankar K, Nagesh R. Estimation of long-term prestress losses
of box-girder bridge span using the prestress loss models in Indian and International codes of practices and comparison with field
measurements. The Indian Concr J. 2015;89(5):59-68.
76. Garber D, Gallardo J, Deschenes D, Bayrak O. Experimental investigation of prestress losses in full-scale bridge girders. ACI Struct J.
2015;112(5):553-564. http://www.concrete.org/Publications/InternationalConcreteAbstractsPortal.aspx?m=details&i=51687909
77. Barr PJ, Kukay BM, Halling MW. Comparison of prestress losses for a prestress concrete bridge made with high-performance concrete.
J Bridge Eng. 2008;13(5):468-475.
78. Anderson TC, Houdeshell DM, Gamble WL. Construction and long-term behavior of 1/8th scale prestressed concrete bridge components,
Vol. 384. Champaign, IL: University of Illinois; 1972.
79. Brewe JE, Myers JJ. High-strength self-consolidating concrete girders subjected to elevated compressive fiber stresses, part 1: Prestress
loss and camber behavior. PCI J. 2010;55(4):59-77. http://www.pci.org/pci_journal-2010-fall-8/
80. Wang WW, Dai JG, Li G, Huang CK. Long-term behavior of prestressed old-new concrete composite beams. 2011;16(April):275-285.
81. Remennikov AM, Kaewunruen S. Determination of prestressing force in railway concrete sleepers using dynamic relaxation technique.
J Perform Constr Facil. 2015;29(5):04014134. http://ascelibrary.org/doi/10.1061/{%}28ASCE{%}29CF.1943-5509.0000634
82. Mckeeman I, Fusiek G, Perry M, et al. First-time demonstration of measuring concrete prestress levels with metal packaged fibre optic
sensors. Smart Mater Struct. 2016;25(9):095051. https://doi.org/10.1088/0964-1726/25/9/095051http://stacks.iop.org/0964-1726/25/i=9/
a=095051?key=crossref.e4356a09af82e79b8c4d2f1c57cd6f50
83. Idriss RL, Liang Z. Monitoring an interstate highway bridge with a built-in fiber-optic sensor system. Proceedings of SPIE, Vol. 6529;
2007:877-878. http://www.scopus.com/inward/record.url?eid=2-s2.0-56749169355&partnerID=40&md5=50e30d7f771118c2386be8f27
b6e171a
84. Dwairi HM, Wagner MC, Kowalsky MJ, Zia P. Behavior of instrumented prestressed high performance concrete bridge girders. Constr
Build Mater. 2010;24(11):2294-2311. http://linkinghub.elsevier.com/retrieve/pii/S0950061810001467
85. Saiidi M, Shields J, O'Connor D, Hutchens E. Variation of prestress force in a prestressed concrete bridge during the first 30 months. PCI
J. 1996;96.
86. Dolatabad YA, Maghsudi AA. Monitoring and theoretical losses of post-tensioned indeterminate I-beams. Mag Concr Res.
2014;66(22):1129-1144. http://www.icevirtuallibrary.com/doi/10.1680/macr.14.00083
87. Gao J, Shi B, Zhang W, Zhu H. Monitoring the stress of the post-tensioning cable using fiber optic distributed strain sensor. Meas.
2006;39(5):420-428. http://linkinghub.elsevier.com/retrieve/pii/S026322410500165X
88. Maaskant R, Alavie T, Measures RM, Tadros G, Rizkalla SH, Guha-Thakurta A. Fiber-optic Bragg grating sensors for bridge monitoring.
Cem Concr Compos. 1997;19(1):21-33.
89. Kim ST, Park Y, Park SY, Cho K, Cho JR. A sensor-type PC strand with an embedded FBG sensor for monitoring prestress forces. Sensors
(Switzerland). 2015;15(1):1060-1070.
90. Zhou ZHI, He J, Chen G, Ou J. A smart steel strand for the evaluation of prestress loss distribution in
post-tensioned concrete structures. J Intell Mater Syst Struct. 2009;20(16):1901-1912. http://10.0.4.153/1045389X
09347021{%}5Cnhttp://search.ebscohost.com/login.aspx?direct=true&db=bth&AN=45103036&site=ehost-live
91. Tabatabai H, Dickson TJ. Structural evaluation of a 34-year-old precast post-tensioned concrete girder. PCI J. 1993;38(5):50-63.
92. Eder RW, Miller RA, Baseheart TM, Swanson JA. Testing of two 50-year-old precast post-tensioned concrete bridge girders. PCI J.
2005;50(3):90-95.
93. Halsey JT, Miller R. Destructive testing of two 40-year-old prestressed concrete bridge beams. PCI J. 1996;41(5):84-93.
94. Labia Y, Saiidi MS, Douglas B. Full-scale testing and analysis of 20-year-old pretensioned concrete box girders. ACI Struct J.
1997;94(5):471-482.
95. Czaderski C, Motavalli M. Determining the remaining tendon force of a large-scale, 38-year-old prestressed concrete bridge girder. PCI J.
2006;51(4):56-68.
ABDEL-JABER AND GLISIC 27 of 27

96. Abdel-Jaber H, Glisic B. A method for the on-site determination of prestressing forces using long-gauge fiber optic strain sensors. Smart
Mater Struct. 2014;23(7):075004. http://stacks.iop.org/0964-1726/23/i=7/a=075004?key=crossref.6e94223489fb2b6f3fccd49be0b881e7

How to cite this article: Abdel-Jaber H, Glisic B. Monitoring of prestressing forces in prestressed concrete
structures—An overview. Struct Control Health Monit. 2019;e2374. https://doi.org/10.1002/stc.2374

You might also like