Rocha 2019

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Nondestructive Evaluation (2019) 38:8

https://doi.org/10.1007/s10921-018-0546-5

Detection of Delaminations in Sunlight-Unexposed Concrete Elements


of Bridges Using Infrared Thermography
Joaquin Humberto Aquino Rocha1 · Yêda Vieira Póvoas1 · Cynthia Firmino Santos1

Received: 9 January 2018 / Accepted: 28 November 2018 / Published online: 1 December 2018
© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
Infrared thermography is a non-destructive test used in the inspection of structures. However, its use for bridge inspection is
still under development and in many cases its application may be limited to bridge sections exposed to direct sunlight. This
study aims to evaluate infrared thermography test for the detection of delaminations in different types of concrete not exposed
directly to sunlight. The experimental methodology uses two concrete specimens with water/cement ratios (w/c) equal to 0.5
and 0.6, through the insertion of polystyrene plates of different thicknesses (3, 6 and 12 mm) and depths (25, 50 and 75 mm),
in order to simulate defects within the concrete and evaluate the capacity of infrared thermography to detect them. The results
show that detection is possible, but limited to short periods of time. In relation to the concrete quality, defects were more easily
detected in the test specimen with lower w/c ratio. It can be said that the nearer to the surface and larger the delamination, the
easier it is to detect. Also, the better the quality of the concrete, the more effective the technique becomes.

Keywords Infrared thermography · Passive thermal imaging · Concrete · Bridge inspection

1 Introduction information, which may vary according to the inspector’s


experience, and may not detect delaminations that are hid-
Bridges are structures vulnerable to deterioration, especially den within the concrete [11]. The hammer and chain drag
the elements of their superstructure, as they must support methods require qualified inspectors and direct contact with
traffic loads and resist aggressive environmental conditions the structure and are very time-consuming [12]. The destruc-
[1, 2]. Corrosion is the principal mechanism of bridge deteri- tive tests can provide detailed information, but only at a single
oration; the formation of iron oxide leads to the expansion of point. In many cases, in addition to causing the disruption of
steel reinforcements. Corrosion generates tensile stress in the service, they can adversely affect the properties of the struc-
concrete, causing splits, cracks, and loosening. It degrades ture [4, 13].
the functionality of the structure and, in many cases, threatens Several non-destructive tests have been developed for
the safety of users [3–6]. Detection of corrosion at its early bridge inspection [14], and they include impact-echo,
stages may reduce and prevent major consequences; how- infrared thermography, ground penetrating radar, and ultra-
ever, there are long intervals between bridge maintenance and sonic test [11]. Several authors have investigated the advan-
inspection programs, which can hamper the timely detection tages and limitations of these methods by laboratory and field
of pathological manifestations in the structures [7]. tests to evaluate their practicality and applicability [15–17].
Bridge inspection is normally performed through tradi- However, many of these techniques are labor-intensive and
tional methods such as visual inspection, hammering, and require the inspector to be close to the structure, which may
chain dragging, and sometimes, destructive tests, such as the require interdiction of the bridge to perform the tests [4, 12].
extraction of concrete cores, are carried out [8–10]. However, These problems can be overcome using IRT because the
visual inspection only provides qualitative and subjective method can be used at a distance. It does not require direct
contact with the structure, removing the need to block traffic,
B Joaquin Humberto Aquino Rocha and it can be used to analyze large areas in real-time. More-
jhar_pec@poli.br over, it requires simple equipment, only a thermographic
camera in most cases [18, 19]. In addition, much progress has
1 Universidade de Pernambuco, R. Benfica 455, Recife, PE been made in recent years in the field of reinforced concrete
CEP 50720-001, Brazil

123
8 Page 2 of 12 Journal of Nondestructive Evaluation (2019) 38:8

bridge inspection because of the continuous development of is because delaminations interrupt the transfer of heat, pro-
technologies and tools that automate the processes of inspec- ducing temperature differentials between the surface of the
tion and data analysis, making them more efficient [2, 20–24]. intact concrete and the defective concrete [7]. The heating
The IRT test is performed by measuring the infrared of concrete through solar radiation produces greater thermal
radiation emitted from the surface of an object using a contrasts than that simply caused by convection, as in the
thermographic camera and subsequently representing it as case of sections not exposed to sunlight [6].
a thermal image. Defects such as delaminations and voids Two modes of IRT—active mode and passive mode—can
within the concrete elements of bridges affect the heat trans- be used for bridge inspection, but the latter is usually adopted.
fer of the concrete and are manifested as thermal contrasts In passive thermography, the structure is under sunlight and
on the surface, which can be detected in the thermal images. the defective areas present a different temperature than the
These thermal differences allow for the identification of surroundings [30]. This undoubtedly limits the test to only a
damaged areas at risk of detachment, and they can help to few sections of the bridge, such as the superstructure, which
prioritize and schedule detailed inspections or, in more sig- are in direct contact with solar radiation. Moreover, in the
nificant cases, repairs [6, 25, 26]. sections unexposed to sunlight, heat is transferred between
Despite the advantages and applicability of the test, ther- the concrete and the environment, and thus, thermal gradi-
mography has some limitations: it cannot detect the depth ents also occur between the damaged and intact areas [12];
of the defects in the concrete (passive thermography). The however, the thermal gradient is reduced because the heat
results are highly sensitive to the onsite environmental condi- transfer depends only on convection, unlike in the sections
tions, and most studies in the literature were performed under directly exposed to the sun.
conditions of constant sun exposure throughout the day [5, In the literature, there are several studies on the detection
8, 9, 12] or with external heat sources (active thermography) of delaminations in reinforced concrete bridges with IRT.
[27, 28]. Considering this, the objective of this study is to Some investigations were carried out through direct inspec-
evaluate the ability of the IRT test in detecting delaminations tion of bridges [4, 29, 31] and others by creating artificial
in concrete elements not exposed to sunlight, while consid- defects of different sizes and depths in the concretes to be
ering different qualities of concrete and defects located at detected [1, 12, 16, 32].
different depths. The study involves an experiment on the Several investigations have shown that the size of the
detectability of defects, and the effects of ambient temper- delamination defects affects the depth up to which they can
ature and relative humidity are considered. The appropriate be detected [4, 8, 16]. Washer, Fenwick, and Bolleni [26]
conditions and effective schedules for the inspection of these found defects of 30 × 30 cm at a depth of 7.6 cm. Kee et al.
defects are explored in order to reduce the uncertainties of [4] detected defects of 61 × 61 cm at a depth of 6.35 cm.
the IRT method and expand its applications to different con- Hiasa et al. [8] detected delaminations of 10.2 × 10.2 cm
ditions. at 2.54 cm depth at different driving speeds. In more recent
works, Tran et al. [28] found defects of 10 × 10 cm and 3 ×
3 cm at depths of 3 cm and 1 cm, respectively. Hiasa et al.
2 Bibliographic Review [32] studied delaminations of 10.2 × 10.2 cm at depths of
1.27 and 2.54 cm. These investigations show that the larger
Radiation, conduction, and convection are the principal and more superficial the defect, the more detectable it is using
mechanisms by which heat can be transferred to concretes. IRT. Based on the results of several authors, Cotič et al. [33]
Radiation is defined as the transfer of heat by means of elec- proposed a relationship involving the depth of defect (z) and
tromagnetic waves [6]. Conduction is the ability of a material dimension of defect (D), concluding that when z/D is less
to allow heat to flow within its interior during a specific than or equal to 0.9, the defect can be detected. Additionally,
period [1]. Convection is the transfer of heat from one place Farrag et al. [1] proposed a similar relationship involving the
to another through a moving fluid. Although radiant energy short dimension of the defect (r) and depth of the defect from
is the parameter that can be measured by a thermographic a surface (D); if r/D is greater than 1, detection is possible.
camera, the thermal conductivity inside the concrete and the Although the estimation of defect depth may be a limitation
convection of heat from the air to the concrete influence what of IRT, methods have been proposed for estimating this value
appears in the thermograms [29]. using active thermography [34].
In bridge elements that are exposed to direct sunlight, the
concrete absorbs heat transmitted by the thermal radiation
of the sun. A convective heat transfer between the envi- 3 Materials and Methods
ronment and the concrete also occurs. These mechanisms
result in thermal gradients on the surface of the concrete To evaluate the ability of thermography to detect delami-
as a result of internal defects such as delaminations. This nations in concretes, two 50 × 50 cm specimens with a

123
Journal of Nondestructive Evaluation (2019) 38:8 Page 3 of 12 8

Table 1 Specifications of
Delamination Concrete specimen 1 (CS-1) Concrete specimen 2 (CS-2)
simulated delaminations
w/c  0.6 w/c  0.5

Depth (mm) Thickness (mm) Depth (mm) Thickness (mm)

D1 25 3 25 3
D2 25 6 25 6
D3 25 12 25 12
D4 50 3 50 3
D5 50 6 50 6
D6 50 12 50 12
D7 50 3 75 3
D8 50 6 75 6
D9 50 12 75 12

10 cm

10 cm

10 cm
5 cm

5 cm

5 cm

5 cm
thickness of 10 cm were molded. The size of the specimens
were determined so as to represent an exposed surface of
the concretes from which surface defects could be observed.
Each test body was created with different water/cement (w/c) 5 cm
ratios, 0.6 and 0.5, to analyze the influence of this charac-
teristic on detection. The proportion of cement:gravel:sand 10 cm
used was 1:2.8:2.6, and the cement content of 1 m3 of the
concrete mixture was 360 kg. The cement used was CPII Z- 5 cm
32, which corresponds to the Pozzolan-modified Portland of
ASTM C 595 [35]. The gravel had a maximum diameter of 10 cm 50 cm
19 mm, and the grain size distribution for both gravel and
sand were in accordance with the limits for concrete aggre- 5 cm
gates indicated in the standard NBR 7211 [36]. To evaluate
10 cm
the technological limitations of detection as regards the depth
and thickness of defects in concrete bridges, delaminations
5 cm
were simulated inside the concrete; nine polystyrene plates
of 10 × 10 cm with different thicknesses were used for each 50 cm
test specimen and were at various depths and positions, as
specified in Table 1 and Fig. 1. Fig. 1 Position of the delaminations in the 10 cm thick concrete speci-
The concreting of the specimens was carried out in four men
layers of 2.5 cm. Pieces of 7.5 cm, 5.0, and 2.5 depths were
placed in the first, second, and third, layers, respectively.
The depth of the polystyrene sheets was controlled by rulers, mined in every thermogram during the monitoring period.
which were placed at ends of the plates and removed after The infrared camera and the optics were kept away from
completion of the concreting. direct sunlight as well as other sources of heat to measure
An FLIR E60 camera was positioned 1 m away from the temperature more accurately.
specimens as shown in Fig. 2. The camera has a thermal Ambient temperature (°C) and relative humidity (%) data,
sensitivity of less than 0.05 °C, a resolution of 320 × 240 which was automatically determined by the camera, was also
pixels and can handle a temperature range of − 20 °C to supplied to compensate for the effects of several different
650 °C. The infrared camera was calibrated by the manufac- radiation sources. These parameters were also collected to
turer 2 months before the tests following recommendations in observe their influence on the thermal behavior of the defects.
the user’s manual, and calibrations must be done every year The test specimens were molded at the University of
[37]. The thermograms were captured at intervals of 30 min Pernambuco, Brazil, following the specifications and rec-
over a period of 14 h from 7:00 a.m. to 9:00 p.m. to analyze the ommendations of NBR 12655 [38]. During the process, four
behavior of the concrete surface throughout the day and deter- cylindrical specimens of 10 cm diameter and 20 cm height
mine the best time to inspect for delaminations. Parameters were also molded for each w/c ratio using the procedure
such as emissivity and reflected temperature were deter- established in NBR 5739 to determine the 28-day compres-
sive strength of the concrete [39]. Table 2 shows the results of

123
8 Page 4 of 12 Journal of Nondestructive Evaluation (2019) 38:8

Fig. 2 Position of the camera to perform the test: a CS-1 and b CS-2

Table 2 Compressive strength of cylindrical concrete specimens for the test did not allow for direct solar radiation (Fig. 3a).
Cylindrical concrete 28-day compressive strength This section can represent parts of the bridge not exposed to
specimens (MPa) the sun, such as the substructure. Here heat transfer occurs
mostly by convection. It also occurs by heat radiation from
w/c  0.6 w/c  0.5
surrounding structures (nearby buildings and the floor) and
CCS-1 24.6 30.6 the atmosphere and by reflection of solar radiation; these
CCS-2 23.6 33.2 contribute approximately 30%–40% of the total heat trans-
CCS-3 27.3 31.5 fer and varies according to the date, hours of the day, and
CCS-4 25.6 33.2 atmospheric conditions. Because analysis is restricted to the
Average 25.3 32.1 surface of the exposed concrete, a wooden scaffolding was
Standard deviation 1.58 1.29 maintained to reduce the edge effects.
Average density 2198 2394 Figure 3b shows the thermogram of the test bodies; the
(kg/m3 ) observable defects are shown as a thermal contrast between
the area of the intact concrete and that of the delaminations. In
the thermogram, the yellow and blue colors represent warm
the compressive strength tests on the cylindrical specimens and cool areas, respectively. In addition to the qualitative
(CS). information of the thermograms, the thermal contrast, T,
After 28 days of curing in a wet chamber, the specimens defined by Eq. (1), was used for the detection of defects and
with delaminations were taken out of the laboratory. The free their subsequent analysis.
moisture contents in the concrete specimens CS-1 and CS-
2 were 3.97% and 3.18%, respectively. The section chosen T  TD − TC . (1)

Fig. 3 Position of concrete


specimens: a digital
photography; b thermogram

123
Journal of Nondestructive Evaluation (2019) 38:8 Page 5 of 12 8

Here, TD is the surface temperature of the delamination or


void, and TC is the temperature of the intact concrete.
The contrast was calculated for each delamination by
considering an area of 6 × 6 pixels within the delamination-
affected area and within the intact concrete area, as shown
in Fig. 3b. The intact concrete point was the same for all
nine plates and was chosen within the area outside the influ-
ence of the defects and edges of the test body. This process
allows a quantitative analysis of the results to be performed
to determine the variation in thermal contrasts throughout the
duration of monitoring and to identify adequate periods of
detection; it would also help determine the influence of the
ambient temperature and relative humidity on the thermal
contrasts.
The monitoring spanned 20 days and was conducted in
July 2017 at times of favorable weather, i.e., without rain or Fig. 4 Behaviors of the delaminations and the intact concrete tempera-
cloudy conditions, and during the test times, the specimens tures for CS-1
were not wet. The reproducibility of the data was proved.

4 Results and Discussion

4.1 Thickness and Depth of Delaminations

The results presented in this section are from one day of


testing, making it possible to generalize the typical behavior
of daily monitoring under ideal environmental conditions for
the development of detectable contrasts.
In Figs. 4 and 5, the behaviors of the delaminations and
the intact concrete temperatures in relation to the time of
data collection, as well as the ambient temperature and rel-
ative humidity, are presented. The results are presented as
smoothed curves for better presentation. The locations of the
CS-1 delamination (w/c ratio  0.6) are shown in Fig. 4. Only Fig. 5 Behaviors of the delaminations and the intact concrete tempera-
delaminations located at a depth of 25 mm (D1, D2, and D3) tures for CS-2
were visible during the testing time. Figure 5 presents the
results corresponding to CS-2 (w/c ratio  0.5) for the delam-
inations at depths of 25 mm (D1, D2, and D3) and 50 mm (D4, perature of the intact concrete remained lower than the
D5, and D6). A difference was observed between the detec- temperature of the delaminations in both specimens, indi-
tion of delaminations for both specimens. For CS-1, only cating the development of positive thermal contrasts. As the
33% (3/9) were detected, while for CS-2, 67% (6/9) were ambient temperature decreased in the afternoon, the temper-
detected. This was due to the difference in the w/c ratio; a ature of the delaminations and the concrete also decreased;
higher w/c ratio may lead to higher porosity, decreasing mass however, the relative humidity increased. During the night,
density, which consequently affects diffusivity; this reduces after 6:00 p.m., a reverse of the behavior observed during
the visible gradients observed with the thermographic cam- the day occurred: the temperature of the concrete was higher
era. [40]. than the temperature of the delaminations, creating negative
For both specimens, with increase in the ambient temper- thermal contrasts.
ature and decrease in the relative humidity, the temperature In both specimens, the maximum temperatures of the
of the concrete delaminations increased. The highest values delaminations occurred at approximately 12:30 p.m., day-
for delaminations and concrete temperatures were recorded time, when the ambient temperature was highest, and after
when the ambient temperature and relative humidity were which the cooling period began. However, the largest ther-
maximum and minimum, respectively, at approximately mal contrasts were produced at 10:30 a.m., as seen in Figs. 6
between 1:00 and 2:00 p.m. Throughout this time, the tem- and 7 in Chapter 4.3.

123
8 Page 6 of 12 Journal of Nondestructive Evaluation (2019) 38:8

is reduced and the temperature of the concrete, which was


heated during the day, tends to equilibrate with the ambient
temperature, avoiding the development of thermal contrasts,
as described by Washer [6]. In this process, relative humidity,
which is high at night, also saturates the concrete; therefore,
the concrete has a lower temperature due to the thermal inertia
of the water, which absorbs more of thermal energy without
a significant alteration of the thermodynamic state.
Considering the thickness of the delamination, the delami-
nations of greater thickness became hotter during the day and
cooler at night when compared to other less-thick delam-
Fig. 6 Thermal contrasts between the delamination and concrete for inations at the same depth, even when the differences in
CS-1 thickness were minimal. The 12 mm thick delaminations
were visible in D3 (25 mm deep) in CS-1 and in D3 and
D6 (25 mm and 50 mm deep, respectively) in CS-2.
The thickness of the delamination influenced the results
because the delamination (polystyrene) allowed only a very
limited transfer of heat (thermal conductivity of approxi-
mately 0.028 W/m·ºC) to the concrete. A greater thickness
implies a greater difficulty in transferring heat to the con-
crete below the delamination, and thus, more heat is retained
in the volume above; therefore, D3 was hotter than D2 and
D1 in CS-1 throughout the day. In CS-2, D3 and D6 were the
hottest for depths of 25 mm and 50 mm, respectively.
At night, the opposite occurred. Radiation exchange con-
Fig. 7 Thermal contrasts between the delamination and concrete for
tributed close to 30% of the total heat transfer, and the
CS-2
delaminations became cooler than the surrounding concrete.
This was mainly due to the temperature balance of the heated
This behavior can be explained by the heating of the con- concrete with low ambient temperature during the day and to
crete. During the day, the ambient temperature increased and a lesser extent, to the reflected temperature of other objects
reached the surface of the concrete, heating it through con- near the concrete. In this case, the areas above the more
vection and the reflection of solar radiation on other objects. superficial delaminations reached equilibrium with the envi-
The heat on the surface was transferred through the concrete, ronment more quickly because they were less thick than
heating it. However, this transfer was reduced when the heat the intact concrete and the deeper delaminations, creating
reached the delamination, which had a thermal conductivity negative temperature contrasts. Like during the day, delami-
different from that of the concrete. The closer the delamina- nations were observable at the 25 mm depth in CS-1 and up
tion was to the surface, the faster the concrete above it heated to 50 mm in CS-2.
up due to its smaller volume. For deeper delaminations, a Although the ambient temperature was the main source of
larger volume would be heated by the same amount of heat thermal gradients on the concrete surface, the relative humid-
transferred from the environment. This is why the tempera- ity also played a role in the thermal contrast. For lower values
tures of the more superficial delaminations at CS-1 (25 mm of relative humidity, greater thermal contrasts developed, and
deep) and CS-2 (25 mm and 50 mm deep) were higher, this occurred approximately from noon to 2:00 p.m. On the
and therefore, the radiations emitted from the surfaces above other hand, when the relative humidity values were higher,
them were higher and were captured by the camera as ther- the thermal contrasts were smaller, and this occurred during
mal contrasts on the surface. The deeper delaminations may the night from 7:00 p.m. to around 9:00 p.m. It was observed
have minimal radiation difference that cannot be captured that the relative humidity influenced the heat transfer pro-
and visualized in thermograms. In addition, the equipment cess in the concrete. As expected, the heat flow was faster
may also influence the detectability of defects. in a humid atmosphere than in a dry atmosphere, increas-
In both specimens, during the night hours, the temperature ing the thermal gradient between the exterior surfaces and
of the defects and the ambient temperature tended toward the concrete core [26, 28]; however, this occurred mainly at
equilibrium, which was more noticeable at 9:00 p.m., as night when the condensation in the concrete improved the
shown in Figs. 4 and 5. This is mainly due to the cooling cycle convective heat transfer, which was low due to the small dif-
that occurs during the night, where the ambient temperature ference between the ambient temperature and the concrete.

123
Journal of Nondestructive Evaluation (2019) 38:8 Page 7 of 12 8

On the other hand, during the afternoon hours, evaporation in the interior of the concrete, interrupting the formation of
occurred due to the high environmental temperatures and the detectable thermal contrasts on the concrete surface.
reflection of solar radiation, preventing the relative humidity The detectability of delamination defects depends on the
from influencing the concrete surface. heat exchange and its temporal behavior, the geometry of
The relative humidity played an important role in the the defects, the thermal resolution of the IR camera, and the
detection of delaminations in both CS-1 and CS-2 because it thermal properties of the material. However, as described in
increased and reduced the moisture content of the concrete the results, visible gradients were observed up to a depth
through the absorption and desorption processes, respec- of 50 mm at CS-2, which had a compressive strength of
tively, which affected several properties of the concrete such 32 MPa. At CS-1, with a compressive strength of 25 MPa,
as diffusivity and thermal conductivity [41–43]. Cement- the results were only visible up to a depth of 25 mm. Com-
based materials generally have type-II sigmoidal sorption pressive strength is a direct consequence of the w/c ratio
isotherms according to the BET classification [44, 45]. and other properties, such as density and porosity [47]. The
Increased relative humidity increased the equilibrium mois- results show that an indirect relationship exists between the
ture content between the concrete and air, improving the compressive strength of the concrete and the maximum depth
thermal conductivity of the concrete. However, when the of detection of the delamination; this corresponds with the
ambient temperature increased, the equilibrium moisture trend found in the results of other investigations [1, 9, 13, 16,
content was lower, indicating that during periods of higher 48], regardless of whether the investigations were performed
temperatures, there were less adsorption and absorption of in conditions of sun exposure or not.
water in the concrete. Moist air penetrates into concretes
faster than dry air, creating a more efficient convective heat 4.3 Time Periods and Inspection Considerations
transfer between the concrete and the environment, resulting
in a thermal gradient, even if the solar energy source is not Figures 6 and 7 show the thermal contrasts of the delam-
present [26]. In this case, there are thermal contrasts at night ination temperature and the concrete for CS-1 and CS-2
with a high degree of relative humidity, whereas during the during daylight hours. For better visualization and presenta-
day, relative humidity is lower. Considering this, it can be tion of the results, the curves have been smoothed. It can be
concluded that the ambient temperature was mostly respon- observed that the more superficial and the greater the thick-
sible for the thermal contrasts in the morning and part of the ness of the delamination, the greater the thermal gradient.
afternoon, whereas the relative humidity contributed mostly Considering the monitoring time, the behavior of delami-
to the process at night by increasing the moisture content of nations can be divided into four distinct periods. The first
the concrete and, therefore, the rate of heat transfer. occurred during the early morning hours, when the gradient
was practically zero, and delaminations were not visible. The
4.2 Water/Cement Ratio second occurred between 10:00 a.m. and 2:00 p.m., when the
gradient reached its maximum, and the delaminations were
In both specimens, defects at a depth of 25 mm were completely visible. During the third period, between 3:00
detectable throughout most of the day because they reacted and 6:00 p.m., the thermal gradient had a positive-to-negative
rapidly to changing conditions, becoming hotter than the sur- inversion; the detection of delaminations became ineffective
rounding concrete, which in this case was thin; therefore, the in this period due to the poor visualization of defects in the
heat was concentrated in a small area and rapid heating (and thermograms. In CS-1, the thermal gradient of D3 was the
rapid cooling at night) was provoked. Delaminations at a first to go below zero and become negative, at approximately
depth of 50 mm were only detected in CS-2, and delamina- 5:00 p.m., and those of D1 and D2 at around 6:00 p.m. In
tions at a depth of 75 mm were not detected at all during the CS-2, the thermal gradient of D3 went below zero to become
test. negative at 5:00 p.m., and those of the other delaminations at
The w/c ratio influenced the detection of defects with IRT around 6:00 and 7:00 p.m., with the exception of D4, whose
since most delaminations detected were in CS-2, which had thermal gradient reached zero at 5: 00 p.m. and remained
a w/c ratio of 0.5, compared to CS-1, which had a w/c ratio of there. Finally, the fourth period occurred at night, after 7:00
0.6. As Al-Hadharmi et al. [46] and Farrag et al. [1] reported, p.m., where a negative contrast was observed, and the delam-
this behavior is due to the porosity and density of the concrete inations were visible again, although to a lesser extent and
(Table 2): Denser concrete has higher thermal conductivity, with lower values than the thermal gradients during the day.
and therefore, more detailed results appeared in the thermo- The thermal contrast developed during the time of the
grams. This explains the existence of more easily detectable experiment was small, approximately 1 °C. However, to
gradients in CS-2 compared to CS-1, which, due to its com- develop higher values, other sources of thermal excitation
position and lower density, was more porous and had a higher could be used through active IRT, as proposed by other
degree of heterogeneity; these properties impede heat flow authors.

123
8 Page 8 of 12 Journal of Nondestructive Evaluation (2019) 38:8

From Figs. 6 and 7, it can be noted that contrast values


greater than 0.4 °C during the day and 0.2 °C at night allow for
the adequate detection of the delamination. Although D4788-
03 [29] minimally establishes 0.5 °C of thermal differential,
and Farrag, Yehia, and Qadduomi [1] consider values above
0.8 °C; moreover, it has been found that detection can be
performed at lower values, with contrasts perceived between
0.2 °C and 0.3 °C, as indicated by Clark, McCann and Forde
Fig. 8 Thermograms at 7:00 a.m.: a CS-1; b CS-2 [18], as well as 0.2 °C, as indicated by Watase et al. [12]. To
quantitatively evaluate the results, the signal-to-noise ratio
of the defects visualized in the figures was calculated. This
quantification is based on the definition of the defective area
and the reference area and is expressed as Eq. (2).
 
 De f μ − Re f μ 
S N R  20 log10 . (2)
Re f σ

Fig. 9 Thermograms at 1:00 p.m.: a CS-1; b CS-2 Here, Defμ is the arithmetic mean of the pixels of the defec-
tive area, Refμ is the arithmetic mean of the pixels of the
reference area, and Refσ is the standard deviation of the pix-
els of the reference area.
The results of the SNR are presented in Table 3. The SNR
values were null for the hours of 7 a.m. and 5 p.m. as there was
no detection in the thermograms; for 1 p.m. the visible defects
in the thermograms presented values greater than 3; however,
by 8 p.m. only some detachments had an SNR greater than
2, indicating that delaminations were detected with greater
Fig. 10 Thermograms at 5:00 p.m.: a CS-1; b CS-2 accuracy during the day.
As revealed in this chapter, the best period to carry out
inspections in sections not exposed to the sun is the sec-
ond period (10 a.m. to 14 p.m.), specifically at noon, since
the highest values of the thermal gradient were developed
and the defects were better defined in the thermograms. The
night period may also be considered, but may have some lim-
itations due to the low values of thermal contrasts. However,
under these conditions, the ambient temperature, which is the
Fig. 11 Thermograms at 8:00 p.m.: a CS-1; b CS-2 main cause of thermal gradients, must be considered. During
the test, it was observed that certain day-night differences
in ambient temperature were required for the development
To better visualize this behavior, Fig. 8 shows the thermo- of thermal gradients in the concrete. A minimum difference
gram captured at 7:00 a.m., representing the behavior during of 5.4 °C allows for the detection of delaminations in the
the first period, where no delaminations were visible. Figure 9 concrete. Table 4 provides the differences in ambient tem-
shows the thermogram captured at 1:00 p.m. in the second perature for the 20 days of the test, in addition to the climatic
period, where the maximum gradients occurred. Figure 10 conditions of the days.
shows the third period when the thermal gradient changed, In this study, the focus is on the results from day 10, where
with a thermogram taken at 5:00 p.m., in which no defects the smallest temperature difference occurred; the days where
were visible. Finally, Fig. 11 shows a thermogram captured at the difference was greater showed higher values of thermal
8:00 p.m., showing the fourth period, where negative thermal contrasts, regardless of the climatic conditions of the day.
gradients were observed. Figures 9 and 11 show that D5 suf- Greater differences in ambient temperature during the day
fered a small displacement from its position during molding allow for the development of greater thermal contrasts, as
in relation to that shown in Fig. 1; however, the depth of each indicated in Washer [6].
delamination was verified during the construction process. It was verified that thermography is strongly influenced
by environmental conditions. When the application is pas-

123
Journal of Nondestructive Evaluation (2019) 38:8 Page 9 of 12 8

Table 3 Signal-to-noise ratio


Time CP1 CP2
values
D1 D2 D3 D1 D2 D3 D4 D5 D6

7:00 0 0 0 0 0 0 0 0 0
13:00 3.51 3.97 4.34 3.88 4.04 4.29 3.09 3.51 3.78
17:00 0 0 0 0 0 0 0 0 0
20:00 1.07 1.92 2.78 1.09 2.01 2.81 0.39 0.84 0.97

Table 4 Differences in ambient


Days of the Relative humidity (%) Temperature (°C) Weather
temperature
test conditions
Maximum Maximum Maximum Minimum Difference

1 45 68 32.8 27.4 6.1 Clear sky


2 43 71 31.5 25.8 5.7 Clear sky
3 41 70 34.3 26.5 7.8 Clear sky
4 41 64 33.6 27.1 6.5 Clear sky
5 42 65 37.6 28.2 9.4 Clear sky
6 45 69 34.5 26.4 8.1 Cloudy night
7 44 71 33.5 27.4 6.1 Clear sky
8 48 78 33.2 26.5 6.7 Clear sky
9 47 72 36.5 28.1 8.4 Clear sky
10 43 62 31.4 26.2 5.2 Clear sky
11 41 68 34.3 26.2 8.1 Clear sky
12 45 72 32.9 25.8 7.1 Clear sky
13 41 69 35.6 27.1 8.5 Clear sky
14 42 64 37.1 28.4 8.7 Cloudy night
15 44 66 35.2 27.3 7.9 Clear sky
16 41 66 32.1 26.5 5.6 Cloudy day
17 45 68 31.8 25.6 6.2 Cloudy day
18 46 70 33.8 26.9 6.9 Clear sky
19 48 66 34.7 27.3 7.4 Clear sky
20 41 67 32.3 26.7 5.6 Cloudy night

sive—as most frequently used for bridge inspections—and Delaminations of much smaller areas would not probably be
the concrete elements are unexposed to direct sunlight, the accurately displayed in the thermograms.
ability to obtain thermograms is affected [5, 6]. However, The detection of defects (delaminations and voids) using
it must be noted that the site chosen for the tests, i.e., the IRT on concrete surfaces involves observing the differences
municipality of Recife-PE, presents adequate environmental in radiation emitted by the surface of the structure, which
conditions for the development of gradients in the concrete. can provide information on the existence of potential defects;
Here, IRT made the delaminations visible both during the however, details about the nature of such defects or irregu-
day and night. larities are limited. In the case of bridges, delaminations and
As investigated, the detection of the depth and size of voids can be linked to corrosion products or building defects.
the delaminations represents a limitation of IRT [7]. In the The existence of temperature gradients is an important ref-
present study, delaminations of the same size were used at dif- erence for the detection of defects in concrete with a high
ferent depths and thicknesses. The maximum depth reached confidence level.
by the thermogram was 50 mm, for a 3 mm thick delami- The results obtained are corroborated by those of other
nation, represented by CS-2 D4. It is important to consider non-destructive techniques, such as ground penetrating radar,
that the area of the studied delaminations (10 × 10 cm) con- ultrasonic tests, and impact-echo [2, 4, 9, 49].
stitutes a high percentage of the surface when compared to Undoubtedly, the simulated delaminations in CS-1 and
others in the literature, and the larger the area of the delami- CS-2 provide a model of the behavior of a defect in con-
nation, the more accurate is the detection in the thermograms. cretes, and the thermal contrast found may be much larger

123
8 Page 10 of 12 Journal of Nondestructive Evaluation (2019) 38:8

than that of a real or natural delamination at a similar depth considering the different environmental conditions and char-
[6]. The data and results show that the change in ambient acteristics that the concrete of the inspected structure is
temperature and other environmental parameters results in subjected to. Nonetheless, the detection of visible thermal
the development of thermal gradients in the concrete, which contrasts in the thermograms accurately indicates the pres-
help to detect and locate hidden damages. If reinforcement ence of defects in the concretes.
steel were present above the delaminations, the results within Infrared thermography is a suitable tool for bridge inspec-
the time the defects visualization would vary, since steel has tion because it is more reliable than solely visual inspection.
greater thermal conductivity than concrete; therefore, the It provides more information, and its application is not
heat flow would be faster when passing through the steel restricted to sections of the bridge exposed to the sun; more-
material, but it would be limited in the delamination. On the over, it also applies to the sections without contact with solar
other hand, if reinforcement were present below the delami- radiation. However, many factors, such as the quality of the
nations, the results would not vary from those shown in this concrete and the environmental conditions of the inspection
study, considering that corrosion would be the main cause of site, must be considered.
these defects. Although this technique allows for the location of defects,
The detection of thermal gradient in the specimens does it is unable to detect their depth. Further investigations using
not necessarily indicate that a real or natural delamination other non-destructive tests may be necessary for better char-
would be observable in the same proportions and with the acterization of the defects.
same values. The thermal contrast would likely present dif-
ferent values for the same conditions and the same depth, but Acknowledgements To CAPES (Coordination of Improvement of
Higher Level Personnel) for the financial support.
its appearance in the thermogram would be a good indica-
tion of the presence of a defect. Because the contrast values
under these conditions may be minimal, the camera used
should have a thermal sensitivity lower than 0.1 °C. References
1. Farrag, S., Yehia, S., Qaddoumi, N.: Investigation of mix-variation
5 Conclusions effect on defect-detection ability using infrared thermography as
a nondestructive evaluation technique. J Bridge Eng. 21, 1–15
(2016). https://doi.org/10.1061/(ASCE)BE.1943-5592.0000779
In this research, a quantitative and qualitative experimental 2. Vemuri, S., Atadero, A.: Case study on rapid scanning techniques
study was performed to analyze the detection of delamina- for concrete bridge decks with asphalt overlay: ground-penetrating
tions in concrete unexposed to direct sunlight through IRT. radar and infrared thermography. Pract. Period. Struct. Des.
Const. 22, 1–8 (2017). https://doi.org/10.1061/(ASCE)SC.1943-
It was demonstrated that the technique is adequate, even
5576.0000313
when the ambient temperature is the only heat source. For 3. Nawy, E.G.: Concrete Construction Engineering Handbook, 2nd
analysis, two specimens of different qualities were molded edn. CRC Press, Upper Saddle River (2008). ISBN 978-0-849-
with polystyrene plates inserted as delaminations. Moreover, 37492-0
4. Oh, T., Kee, S., Arndt, R., Popovics, J., Zhu, J.: Comparison of NDT
parameters such as ambient temperature, relative humidity,
methods for assessment of a concrete bridge deck. J. Eng. Mech-
depth, and thickness of delaminations were considered. The ASCE 193, 305–314 (2013). https://doi.org/10.1061/(ASCE)EM.
results present the general behavior of delaminations and 1943-7889.0000441
their influence on the surface of the concrete with regard to 5. Washer, G., Fenwick, R., Bolleni, N.: Effects of solar loading
on infrared imaging of subsurface features in concrete. J. Bridge
the development of detectable thermal contrasts. Eng. 15, 384–390 (2010). https://doi.org/10.1061/(ASCE)BE.
Inspection with IRT can be effective in detecting delami- 1943-5592.0000117
nations or voids. The closer to the surface the defects are, the 6. Washer, G.: Advances in the use of thermographic imaging for the
easier it is to detect them. However, without exposure of the condition assessment of bridges. Bridge Struct. 8, 81–90 (2012).
https://doi.org/10.3233/BRS-2012-0041
structure to the sun, the inspection time is reduced to short 7. Hiasa S (2016) Investigation of infrared thermography for sub-
periods. surface damage detection of concrete structures. Dissertation,
The variation in the w/c ratio and its impacts on the den- University of Central Florida
sity, thermal conductivity, and heat capacity of the concrete 8. Hiasa, S., Catbas, F.N., Matsumoto, M., Mitani, K.: Monitoring
concrete bridge decks using infrared thermography with high speed
influence the detection of defects. In the detection of delam- vehicle. Struct. Monit. Maint. 3, 277–296 (2016)
inations, the concrete with a lower w/c ratio showed the 9. Kee, S., Oh, T., Popovics, J., Arndt, R., Zhu, J.: Nondestructive
best results due to its higher density. The results show that bridge deck testing with air-coupled impact-echo and infrared ther-
the higher the concrete quality, the more effective the IRT mography. J. Bridge Eng. 17, 928–939 (2012). https://doi.org/10.
1061/(ASCE)BE.1943-5592.0000350
is. 10. Vemuri S (2016) Evaluation of rapid scanning techniques for
The experimental results can probably be very differ- inspecting concrete bridge decks with asphalt overlay. Disserta-
ent from results obtained from real structures inspection, tion, Colorado State University

123
Journal of Nondestructive Evaluation (2019) 38:8 Page 11 of 12 8

11. Rehman, S.K., Ibrahim, Z., Memon, S.A., Jameel, M.: Nondestruc- detection in concrete structures by active thermal imaging. Sen-
tive test methods for concrete bridges: a review. Constr. Build. sors 17, 1–18 (2017). https://doi.org/10.3390/s17081718
Mater. 107, 58–86 (2016). https://doi.org/10.1016/j.conbuildmat. 29. Vaghefi, K., Ahlborn, T., Harris, D.K., Brooks, C.: Combined imag-
2015.12.011 ing technologies for concrete bridge deck condition assessment.
12. Watase, A., Birgul, R., Hiasa, S., Matsumoto, M., Mitani, K., J. Perform. Constr. Fac. 29, 1–8 (2015). https://doi.org/10.1061/
Catbas, F.N.: Practical identification of favorable time windows (ASCE)CF.1943-5509.0000465
for infrared thermography for concrete bridge evaluation. Con- 30. ASTM: D4788–03: standard test method for detecting delam-
str. Build. Mater. 101, 1016–1030 (2015). https://doi.org/10.1016/ inations in bridge decks using infrared thermography. ASTM
j.conbuildmat.2015.10.156 International, West Conshohocken (2013)
13. Alfredo-Cruz, R.A., Quintero-Ortiz, L.A., Galán-Pinilla, C.A., 31. Hiasa, S., Catbas, N., Matsumoto, M., Mitani, K.: Considerations
Espinosa-García, E.J.: Evaluación de técnicas no destructivas en and issues in the utilization of infrared thermography for concrete
elementos de concreto para puentes. Rev. Fac. Ing. 24, 83–96 bridge inspection at normal driving speeds. J. Bridge Eng. 22, 1–11
(2015) (2017). https://doi.org/10.1061/(ASCE)BE.1943-5592.0001124
14. Scott, M., Rezaizadeh, A., Delahaza, A., Santos, C., Moore, M., 32. Hiasa, S., Birgul, R., Matsumoto, M., Catbas, N.: Experimental and
Graybeal, B., Washer, G.: A comparison of nondestructive evalua- numerical studies for suitable infrared thermography implementa-
tion methods for bridge deck assessment. NDT&E Int. 36, 245–255 tion on concrete bridge decks. Measurement 121, 144–159 (2018).
(2003). https://doi.org/10.1016/S0963-8695(02)00061-0 https://doi.org/10.1016/j.measurement.2018.02.019
15. Abudayyeh, O., Abdel-Qader, I., Nabulsi, S., Weber, J.: Using 33. Cotič, P., Kolarič, D., Bokan Bosiljkov, V., Bosiljkov, V., Jagličić,
non-destructive technologies and methods in bridge management Z.: Determination of the applicability and limits of void and
systems. J. Urban Tech. 11, 63–76 (2004). https://doi.org/10.1080/ delamination detection in concrete structures using infrared ther-
1063073042000341989 mography. NDT&E Int. 74, 87–93 (2015). https://doi.org/10.1016/
16. Yehia, S., Abudayyeh, O., Nabulsi, S., Abdelqader, I.: Detection of j.ndteint.2015.05.003
Common Defects in Concrete Bridge Decks Using Nondestructive 34. Vavilov, V.: Pulsed thermal NDT of materials: back to the basics.
Evaluation Techniques. J. Bridge Eng. 12, 215–225 (2007) Nondestruct. Test. Eval. 22, 177–197 (2007). https://doi.org/10.
17. Forde, M.C.: International practice using NDE for the inspection of 1080/10589750701448407
concrete and masonry arch bridges. Bridge Struct. 6, 25–34 (2010). 35. ASTM: C595 / C595 M-17, Standard Specification for Blended
https://doi.org/10.3233/BRS-2010-004 Hydraulic Cements. ASTM International, West Conshohocken
18. Clark, M., Mccann, D., Forde, M.C.: Application of infrared ther- (2017)
mography to the non-destructive testing of concrete and masonry 36. ABNT: NBR 7211: Agregado Para Concreto. ABNT, Rio de
bridges. NDT&E Int. 36, 265–275 (2003). https://doi.org/10.1016/ Janeiro (2009)
S0963-8695(02)00060-9 37. FLIR: User’s Manual FLIR Exx Series, 1st edn. FLIR, Wilsonville
19. Vaghefi, K., Oats, R., Harris, D., Ahlborn, T., Brooks, C., Endsley, (2014)
K., Roussi, C., Shuchman, R., Burns, J., Dobson, R.: Evaluation 38. ABNT: NBR 12655: Concreto de Cimento Portland – Preparo,
of commercially available remote sensors for highway bridge con- controle e recebimento. ABNT, Rio de Janeiro (2015)
dition assessment. J. Bridge Eng. 17, 886–895 (2012). https://doi. 39. ABNT: NBR 5739: Concreto-Ensaio de compressão de corpos-de-
org/10.1061/(ASCE)BE.1943-5592.0000303 prova cilíndricos. ABNT, Rio de Janeiro (2007)
20. Bagavathiappan, S., Lahiri, B., Saravanan, T., Philip, J.: Infrared 40. Marshall, A.L.: The thermal properties of concrete. Build-
thermography for condition monitoring: a review. Infrared Phys. ing Science 7, 167–174 (1972). https://doi.org/10.1016/0007-
Technol. 60, 35–55 (2013). https://doi.org/10.1016/j.infrared. 3628(72)90022-9
2013.03.006 41. Xi, Y., Bažant, Z.P., Jennings, H.M.: Moisture diffusion in cementi-
21. Dabous, S., Yaghi, S., Alkass, S., Moselhi, O.: Concrete bridge deck tious materials Adsorption isotherms. Adv. Cem. Mater. 1, 248–257
condition assessment using IR Thermography and Ground Pene- (1994). https://doi.org/10.1016/1065-7355(94)90033-7
trating Radar technologies. Automat. Constr. 74, 340–354 (2017). 42. Taoukil, D., El bouardi, A., Sick, F., Mimet, A., Ezbakhe, H.,
https://doi.org/10.1016/j.autcon.2017.04.006 Ajzoul, T.: Moisture content influence on the thermal conductivity
22. Ellenberg, A., Kontsos, A., Moon, F., Bartoli, I.: Bridge Deck and diffusivity of wood-concrete composite. Constr. Build. Mater.
delamination identification from unmanned aerial vehicle infrared 48, 104–115 (2013). https://doi.org/10.1016/j.conbuildmat.2013.
imagery. Automat. Constr. 72, 155–165 (2016). https://doi.org/10. 06.067
1016/j.autcon.2016.08.024 43. Gomes, M.G., Flores-Colen, I., Manga, L.M., Soares, A., Brito,
23. Hiasa, S., Birgul, R., Catbas, N.: A data processing methodology J.: The influence of moisture content on the thermal conductivity
for infrared thermography images of concrete bridges. Comput. of external mortars. Constr. Build. Mater. 135, 279–286 (2017).
Struct. 190, 205–218 (2017). https://doi.org/10.1016/j.compstruc. https://doi.org/10.1016/j.conbuildmat.2016.12.166
2017.05.011 44. Abdelhamid, M., Mihoubi, D., Sghaier, J., Bellagi, A.: Water sorp-
24. Rocha, J.H.A., Póvoas, Y.V.: Infrared thermography as a non- tion isotherms and thermodynamic characteristics of hardened
destructive test for the inspection of reinforced concrete bridges: a cement paste and mortar. Transport Porous Med. 113, 283–301
review of the state of the art. Rev. ALCONPAT 7, 200–2014 (2017) (2016). https://doi.org/10.1007/s11242-016-0694-y
25. Manning, D.G., Holt, F.B.: Detecting delaminations in concrete 45. Burgh, J., Foster, S.: Influence of temperature on water vapour
Bridge Decks. Concr. Int. 2, 34–42 (1980) sorption isotherms and kinetics of hardened cement paste and con-
26. Washer, G., Fenwick, R., Bolleni, N.: Development of Hand-held crete. Cem. Concr. Res. 92, 37–55 (2017). https://doi.org/10.1016/
Thermographic Inspection Technologies, 1st edn. MODOT, Jeffer- j.cemconres.2016.11.006
son City (2009) 46. Al-hadharmi, L.M., Maslehuddin, M., Shameem, M., Ali, M.:
27. Baek, S., Xue, W., Feng, M.Q.: Nondestructive corrosion detec- Assessing concrete density using infrared thermographic (IRT)
tion in RC through integrated heat induction and IR thermography. images. Infrared Phys. Technol. 55, 442–448 (2012). https://doi.
J. Nondestruct Eval. 31, 181–190 (2012). https://doi.org/10.1007/ org/10.1016/j.infrared.2012.04.004
s10921-012-0133-0 47. Mehta, K.P., Monteiro, P.: Concrete: Microstructure, Properties,
28. Tran, Q.H., Han, D., Kang, C., Haldar, A., Huh, J.: Effects of and Materials, 4th edn. McGraw-Hill, New York (2012). ISBN
ambient temperature and relative humidity on subsurface defect 978-0-071-79787-0

123
8 Page 12 of 12 Journal of Nondestructive Evaluation (2019) 38:8

48. Maierhofer, C., Arndt, R., Rollig, M.: Influence of concrete 49. Aggelis, D., Kordatos, E., Soulioti, D., Matikas, T.: Combined use
properties on the detection of voids with impulse-thermography. of thermography and ultrasound for the characterization of sub-
Infrared Phys. Technol. 49, 213–217 (2007). https://doi.org/10. surface cracks in concrete. Constr. Build. Mater. 24, 1888–1897
1016/j.infrared.2006.06.007 (2010). https://doi.org/10.1016/j.conbuildmat.2010.04.014

123

You might also like