You are on page 1of 23

Received: 2 December 2016 Revised: 9 March 2017 Accepted: 22 April 2017

DOI: 10.1002/stc.2042

RESEARCH ARTICLE

Experimental and finite element investigation of temperature


distributions in concrete‐encased steel girders

Sallal R. Abid1 | Faten Mussa2 | Nildem Tayşi2 | Mustafa Özakça2

1
Department of Civil Engineering, Wasit
Summary
University, Wasit, Iraq
2
Department of Civil Engineering,
The structural performance of bridge structures is temporal and is mainly con-
Gaziantep University, Gaziantep, Turkey trolled by the types of the applied loads. To continuously observe the structural
performance of bridges, structural health monitoring sensors that include among
Correspondence
Sallal R. Abid, Department of Civil many temperature sensors are used. The impact of nonuniform temperature distri-
Engineering, Wasit University, Wasit, Iraq. butions in bridge girders due to the environment thermal loads has been recog-
Email: sallal@uowasit.edu.iq
nized by former researchers and bridge design codes. To evaluate these and
other effects on the structural behavior of bridge structures, many field and exper-
imental structural health monitoring studies were carried out. However, more
researches are required to investigate the temperature distributions in other girder
configurations. This work is directed to investigate the impact of air temperature
and solar radiation on temperature gradient distributions in concrete‐encased
composite girders. For this purpose, an experimental concrete‐encased steel girder
segment was instrumented with thermocouples and other sensors. The experimen-
tal data recording continued for 6 months during the hot and cold seasons. Further-
more, a thermal finite element (FE) parametric study was conducted to investigate
the effect of the girder size. The test results showed that the vertical and lateral
temperature gradient distributions and the variation of the temperature gradients
with time are controlled by the amount and location of the received solar radia-
tions. The FE analysis showed that the daily temperature variations are higher in
smaller girders, whereas the temperature gradients are smaller than in larger
girders. Moreover, the FE results showed that the thickness of the girder's concrete
members has an important impact on temperature gradients and temperature
distributions.

K EY WO R D S
composite girder, solar radiation, temperature distribution, temperature gradient, thermal strain

1 | INTRODUCTION

As bridges are generally built in open areas and subjected to the daily and seasonally fluctuation of the environment thermal
loads, they are subjected to undesirable deformations and stresses.[1–3] The solar radiation causes the temperature to vary
nonlinearly through the depth and the width of the girders. Such variation leads to additional stresses (equilibrating stresses)
to prevent the nonlinear deformation of the girder cross section. Moreover, the average temperature of the whole structure
changes during the day and night cycles causing longitudinal movements or continuity stresses based on the end conditions.[4]

Struct Control Health Monit. 2017;e2042. wileyonlinelibrary.com/journal/stc Copyright © 2017 John Wiley & Sons, Ltd. 1 of 23
https://doi.org/10.1002/stc.2042
2 of 23 ABID ET AL.

According to previous studies, the stresses that can be referred to as thermal equilibrating stresses can be comparable to those
caused by gravity loads and can result in concrete cracking.[4–6]
Many field studies[7–14] were conducted during the last 30 years to better understand the temperature distributions and their
structural effects on concrete, steel, and composite bridge structures. Generally, field studies are sponsored by the transportation
agencies as parts of complete long‐term structural health monitoring (SHM) projects. Due to their importance, bridges became a
primary focus of SHM techniques and damage detection methods.[15,16] The long‐term SHM systems include the instrumenta-
tion of parts of the bridge, including the superstructure, by many measurement sensors and data acquisition systems. Due to the
high cost of bridge instrumentation and the enormous sensors required for such measurements, smaller size experimental girder
segments were used by many previous studies[17–24] to evaluate the effect of the environmental thermal loads on the temperature
distributions, structural performance, and on the reliability of the obtained data from the SHM systems. Other studies[6,19,25–34]
used analytical, numerical, and statistical techniques for this purpose.
In previous studies,[20,21,33] a full‐scale experimental concrete box‐girder segment was instrumented with 64 thermocouples
and a weather station to analyze the temperature distributions in such superstructures. The thermocouples were distributed in
different locations along the webs and flanges of the girder to evaluate the vertical and lateral temperature gradients. Based
on the analysis of the collected data, design gradients and equations were introduced. Kulprapha and Warnitchai[18] constructed
and tested experimental 8.0 m length continuous prestressed concrete bridge girders. The tested girders have two spans each of
4.0 m. The cross‐sectional dimensions were scaled down with approximately 1/10 to 1/20 scale factors where the section depth
was 300 mm only. Electrical resistance strain gages were used to measure the strain in concrete, reinforcing bars, and
prestressing tendons. Type K thermocouples were used to monitor the temperature distributions over the experimental girders.
Lee and Kalkan[17] measured the temperature variation in a 1.6 m depth and approximately 1.5 m length experimental
prestressed girder segment using 18 thermocouples distributed along the web and the flanges. On the basis of the recorded tem-
perature measurements and thermal loads data, they proposed formulas for maximum vertical and lateral temperature differen-
tials. Wang et al.[22] used experimental model of a box‐girder arch bridge to evaluate the effect of nonlinear temperature
gradients on such bridges. The experimental data of temperature from 1‐day measurements were used to verify finite element
(FE) studies. Xia et al.[23] conducted an experimental work on a simply supported slab specimen to investigate the effect of tem-
perature gradients on the structural vibration characteristics. The slab was 3.0 m span, 0.12 m thickness, and 0.8 m wide and
was instrumented by 11 thermocouples and 14 accelerometers. Following researchers[34] used the experimental data of Xia
et al.[23] to numerically study the impact of environmental thermal loads on the structural frequencies of T‐beam and box girder
bridge structures. Liu et al.[24] carried out 3‐day temperature measurements on an H‐beam steel specimen under solar radiation.
The web depth and flanges width of the specimen were 200 mm, whereas the length of the specimen was 500 mm. The spec-
imen was instrumented with10 temperature sensors.
The current bridge design codes recognize the thermal actions in bridge girders and deck slabs. To overcome such actions,
these codes provide design procedures using temperature gradient models along the depth of the superstructure. Despite that
these codes do not explicitly specify the depth limitations for the use of these gradient, the details of these design gradients have
some restrictions that limit the minimum applicable depth of the section, whereas no values are specified for the upper limit of
the superstructure depth. The AASHTO LRFD Bridge Design Specifications[35] considers the proposed design temperature gra-
dient for members as shallow as 400 mm. The British standard BS 5400‐2[36] recognizes concrete girders from composite ones.
For concrete girders, the design gradient is applicable for sections as shallow as 350 mm; whereas for composite girders, the
design gradient is applicable for sections not shallower than 600 mm. The European standard EN 1991‐1‐5[37] introduces
two procedures to evaluate temperature gradients across the depth of the bridge superstructure. The first assumes a linear tem-
perature difference between the top and bottom surfaces of the section. This procedure does not mention any limitation for the
size of the girder. On the other hand, the second procedure is that of the BS 5400‐2.[36] The New Zealand Bridge Design Manual
SP/M/022[38] introduces a nonlinear gradient model for superstructures with depth not less than 1.4 m. However, it also gives
recommendation for the application of this gradient model to sections shallower than 1.4 m.
The above literature review shows that different size and shape of experimental specimens have been used as experimental
SHM girders or decks to evaluate the effect of solar radiation and air temperature on the distributions of temperature and its
consequent effects on the structural behavior of different types of girders. However, no previous study has investigated temper-
ature gradient distributions in concrete‐encased steel girders. Also, it is obvious in the introduced review that the proposed tem-
perature gradient models are considered applicable for wide range of girder sizes, which is quite questionable.
In this research, a composite concrete‐encased steel girder segment was instrumented with several thermocouples and other
sensors to evaluate the temperature gradient distributions in composite girders. The variation of these gradients with the overall
size of the girder, the depth of the girder, and the thickness of the flanges and the web were numerically studied using a FE
thermal model, which was verified with the experimental temperature measurements. Moreover and as discussed earlier, as
ABID ET AL. 3 of 23

temperature effects are dependent of the time of the day and the year, SHM system was used to observe the long‐term impact of
temperature variation. On the other hand, the reliability of the different data acquired from the SHM acquisition systems is also
affected by this temperature variation. To estimate and correct this effect, studies that use analytical and numerical techniques,
which are verified with field or experimental data, were carried out and are still required. Therefore, the experimental data pre-
sented in this study can also be used as a verification tool for future diagnostic studies.

2 | E X P E R I M E N TA L P RO G R A M

To study the influence of the ambient thermal loads on composite girders, an experimental work was directed using an exper-
imental composite girder segment. The time‐dependent ambient thermal loads include primarily the thermal solar radiation in
addition to the temperature and the speed of the contacting air. In this section, the preparation, geometry, and instrumentation of
the experimental girder segment are explained.

2.1 | The girder segment


A composite girder segment was prepared, instrumented, and cast in this work. The girder segment composed of an I‐shape
steel beam embedded in a concrete section. The steel beam was fabricated of 8 mm thickness plates. The depth of the steel sec-
tion is 400 mm, whereas the width of the flanges is 200 mm. On the other hand, the geometry and dimensions of the encasing
concrete section are shown in Figure 1. The top flange of the girder was reinforced with steel mesh composed of 10 mm diam-
eter steel reinforcing bars spaced at 100 mm in both directions as shown in Figure 2a. The reinforcing steel mesh was placed at
the central depth of the top flange. The girder segment was cast of normal concrete that has an average cube compressive
strength of 42.8 MPa. The mix proportions of the concrete in kg/m3 were (480: 705: 810: 192) (cement: sand: gravel: water).
In addition, 0.5% superplasticizer was used to enhance the consistency of concrete and to facilitate the casting in the formwork.
After preparation and instrumentation, the steel section was embedded in the formwork of the concrete section and the form-
work was sealed except at the top surface. Then after, the reinforcing steel mesh was instrumented and placed in the formwork
of the top flange as shown in Figure 2a. Figure 2b shows the girder segment after concrete casting. The figure also shows the
insulation plates that were used to thermally insulate the section along the longitudinal axis of the girder. The longitudinal axis
was placed along the east‐west direction so that one of the webs is facing the north whereas the other faces the south.

2.2 | Distribution of sensors


Three groups of sensors were used in this study. The first group was the sensors of the ambient thermal loads, which include a
pyranometer to measure the intensity of solar radiation, an anemometer to measure the speed of the wind, and an air temper-
ature sensor. On the other hand, the second group consisted of 15 thermocouples distributed in the concrete and on the steel
section to measure the temperature of the different parts of the girder. All thermocouples were placed at the mid‐length section,

FIGURE 1 The geometry of the composite girder segment


4 of 23 ABID ET AL.

FIGURE 2 The experimental girder segment: (a) the formwork and the distributed sensors of the top flange and (b) the final view of girder and the
insulation boards

TABLE 1 The coordinates of the thermocouples

Thermocouple x (mm) z (mm) Thermocouple x (mm) z (mm) Thermocouple x (mm) z (mm)

TC1 0 500 TC6 200 450 TC11 −50 250


TC2 0 475 TC7 400 450 TC12 50 250
TC3 0 450 TC8 −150 50 TS1 0 425
TC4 −400 450 TC9 150 50 TS2 −4 250
TC5 −200 450 TC10 0 0 TS3 0 25

thus at 250 mm from the insulated edges. The coordinates of the thermocouples are listed in Table 1; in addition, Figure 2a
shows the thermocouples distributed in the top flange. In Table 1, TC refers to the thermocouples embedded in the concrete,
whereas TS refers to the thermocouples that measure the temperature of the steel section. The origin point (0, 0) in Table 1
was considered as the center of the bottom surface of the girder. The third group of sensors consisted of two vibrating wire strain
gages to measure the thermal strains of the concrete. The strain gages were placed on the reinforcing mesh in two perpendicular
directions as shown in Figure 2a.
After concrete casting, all wires from thermocouples and strain gages were grouped and connected to the data acquisition
system, which composed of the data logger, multiplexers, and vibrating wire strain analyzer from Campbell Scientific. The data
from all sensors were collected from the first of July until the end of December 2015 at time intervals of 30 min. However,
during this period, there were some faults that interrupted the data collecting for discrete intervals.

3 | E X P E R I M E N TA L T E M P E R AT U R E S A N D STR A I N S

Three main factors affect the temperature and hence the temperature gradients in bridge girders. These are the daily maximum
and minimum air temperatures, the solar radiation intensity, and the movement of the sun from rise to set. Based on these influ-
ential factors, the distributions of vertical and lateral temperature gradients are different in summer than in winter. These distri-
butions are studied in this section. To optimize the analysis of the experimental records, 10 days from summer and 10 days from
winter were selected to visualize the thermal effects of the environmental thermal loads on the experimental composite girder.
From July 1 to 10 were selected as the summer representative days, whereas from December 21 to 30 were selected to represent
winter. The daily variations of the temperature gradients and thermal strains in addition to the distributions of the maximum
vertical and lateral temperature gradients are discussed in this section for the two periods.
The maximum temperature gradient in this paper refers to the difference between the girder's maximum and minimum tem-
peratures along a specified section. All sections were considered along the position of thermocouples at the mid‐length cross
section. The vertical temperature gradient was calculated from the concrete and steel thermocouples passing through the vertical
centerline of the cross section, hence all thermocouples having zero x‐coordinates as listed in Table 1. These thermocouples,
from top to bottom, are TC1 to TC3, TS1 to TS3, and TC10. The top‐flange lateral temperature gradient was calculated from
ABID ET AL. 5 of 23

the thermocouples distributed horizontally along the centerline of the top flange, thus, TC3 to TC7. On the other hand, the web
lateral temperature gradient was calculated at the central depth of the cross section from thermocouples TC11, TS2, and TC12.

3.1 | Solar radiation and air temperature during the selected periods
Figure 3a shows the hourly variations of air temperature and solar radiation from July 1 to 10, whereas Figure 3b shows these
variations from December 21 to 30. Figure 3a shows that the daily maximum temperatures were close during the first week of
July with an edge of approximately 35 °C, whereas it decreased in July 8 and 9 to less than 32 °C and increased in July 10 to
slightly more than 36 °C. Another notice is that the difference between the daily maximum and minimum was high during the
first 5 days of July, whereas it was low during the following 4 days. The variation of the daily maximum solar radiation was
minimal during this period where values around 1,000 W/m2 were recorded for the 10 days.
Figure 3b shows that solar radiation curves are not smooth during some of the 10 days from December 21 to 30, which
reflects the existence of a variable cloud cover during these days. In general, the daily maximum hourly solar radiation in this
period ranged from approximately 510 W/m2 to approximately 580 W/m2, except in December 29, where it was approximately
300 W/m2. The figure also shows that the daily minimum air temperature during this period was mostly less than 0 °C, whereas
the daily maximum temperature exhibited higher fluctuation.

3.2 | Temperature gradients in July


Figure 4 shows the hourly variation of temperature gradients during the 10 days of July. As shown in Figure 4a, after midnight
and during the early morning hours before sunrise, the vertical temperature gradients become the lowest reaching negative
values, which mean that the temperature of the top surface becomes lower than the temperature of the section interiors. Starting
from sunrise, the vertical gradient starts the ascending phase (rising region). The ascending phase of the gradient is composed of
two parts as shown in Figure 4a. The first has a low slope (slow rising), which occurs during the early day hours from sunrise to

FIGURE 3 Ten‐day variation of the environmental thermal loads: (a) July 1 to 10 and (b) December 21 to 30.
6 of 23 ABID ET AL.

FIGURE 4 Hourly gradients from July 1 to 10: (a) vertical gradient; (b) lateral top‐flange gradients; and (c) lateral web gradients

about 7:00 in the morning, after which the gradient increases sharply (quick rising) until reaching the maxima around 13:00.
The temperature gradient is the highest from midday to about 14:00 and starts the sharp descending phase (dropping region)
after 15:00. The gradient continues to decrease after sunset until starting the next day's loop at the sunrise.
After sunrise, the solar radiation is low and is almost perpendicular to vertical surfaces and thus the radiation reaching the
top surface is low to cause serious increase in the gradient. Therefore, the gradient increases slowly during this period as shown
in the first ascending phase in Figure 4a. After about 2 hr, the altitude of sunrays becomes higher, leading to an effective
worming of the top surface of the girder. This altitude increases and solar radiation increases sharply after 9:00, leading to much
higher vertical temperature gradients, which explains the second ascending phase of the gradient. The sun movement reverses
after midday, and the sunrays altitude starts to decrease showing a significant decrease in the worming budget after 15:00.
ABID ET AL. 7 of 23

Within the sunset, the cooling phase starts by two means. The first is the reradiation of heat from the girder to the surrounding
ambient by long waves, whereas the second is the convection from the surrounding cold air. The cooling phase continues
decreasing the temperature of the top surface and hence decreasing the vertical temperature gradient until the next day's
morning.
The lateral temperature gradients measured at the northern edges (Figure 4b,c) show different behavior from that of
the vertical temperature gradient shown in Figure 4a, whereas those measured at the southern edges show an agreement
with the behavior of the vertical gradient. Lateral temperature gradients at the northern edges exhibit two periods of
rising and dropping during the single day: major rising after sunrise and minor rising before sunset. Mentioning that
in summer, the sun rises from the northeast and sets at the northwest would explain this behavior. During these two
periods, the altitude angle of sunrays is too low. Thus, solar radiation is concentrated on the northern edge of the top
flange and the northern surface of the web leading to the rising of the temperature of these surfaces compared to the
cold interiors, hence, increases the lateral gradients there. As shown in Figure 4a,b, as temperature gradients at the north-
ern surfaces start dropping, those at the south surfaces start rising. This is because the sun moved from the sunrise posi-
tion toward the south.
In general, the recorded values of lateral temperature gradients were much lower than those of vertical gradient in summer.
The recorded maximum lateral temperature gradient along the top slab for the period from July 1 to 10 was 5.9 °C. Similarly,
the maximum lateral temperature gradient across the web was only 4.9 °C, whereas the recorded maximum vertical gradient was
15.4 °C.

3.3 | Temperature gradients in December


It is obvious from the comparison between Figures 4a and 5a that vertical temperature gradients in summer are much higher
than in winter. During the 10 days from December 21 to 30, the recorded maximum vertical temperature gradient was
3.2 °C, whereas along the 10 days of July, it was 15.4 °C. In addition to the values of vertical gradients, another difference
can be distinguished between the behaviors of the gradient variation with time in July and December. It is clear that the slow
rising region is minimal compared to the rapid rising region in summer as shown in Figure 4a, whereas it is significant com-
pared to the rapid rising region in winter as shown in Figure 5a. This is again because of the different inclination angles of
sunrays between summer and winter. In summer, as the sun rises with low altitude angles, it quickly moves toward the zenith
showing small zenith angles (high altitudes) leading to the concentration of solar radiations on the top surface, which rapidly
increases the vertical gradient. On the other hand, in winter, the solar angles keep moving from sunrise to sunset with low alti-
tudes. Hence, the vertical gradients increase slowly.
Comparison between Figures 4b and 5b show that lateral temperature gradients along the top flange are significantly
higher in winter than in summer. The recorded maximum lateral temperature gradient along the first 10 days of July was
5.9 °C, whereas that recorded along the last 10 days of December was 12.4 °C. Another obvious difference between the
behaviors of the two lateral gradients with time is the edge at which the maximum gradient occurred. In summer, and as
explained in the previous section, higher lateral temperature gradients occurred at the north edge because the sun rises
from northeast and sets at northwest. Oppositely, the south edges showed higher lateral temperature gradients in winter,
both along the top flange as shown in Figure 5b and across the web as shown in Figure 5c. This is directly attributed to
sun movement. In winter, the sun rises from southeast moves at low altitude angles to the south of the equatorial plane
until setting at the southwest. Thus, the sun never crosses the equatorial plane in winter. This means that along the
whole day, solar radiations strike the southern vertical edges, leading to a significant increase in the temperature of
theses edges compared to the other points along any horizontal plane, which in turn increase the lateral temperature
gradients.

3.4 | Maximum vertical temperature gradients during the selected periods


Understanding the distribution of the maximum vertical temperature gradient is required in the design of bridge girders. The
maximum vertical temperature gradient during the first 10 days of July occurred in July 25 and was 15.4 °C; whereas during
the last 10 days of December, the maximum was only 3.2 °C and was recorded in December 25.
It is shown in Figure 6 that the distribution is significantly different between the summer gradient (July 25) and that of
winter (December 25). Due to the high concentration of sun rays on the top surface of the girder, the gradient was high there
and across the top flange compared to the lower parts of the girder. Because of the low thermal conductivity of concrete, the
temperature gradient decreased sharply across the top flange. The center of the girder receives only low amounts of solar energy
8 of 23 ABID ET AL.

FIGURE 5 Hourly gradients from December 21 to 30: (a) vertical gradient; (b) lateral top‐flange gradients; and (c) lateral web gradients

during the early morning hours, whereas it is completely shaded during the mid‐day hours at which the solar radiation is
maxima. Similarly, the lower surface of the girder is almost completely shaded along the whole day hours. However, reflected
radiations from the ground led to a slight increase in its temperature compared to thermocouple TS3 (within the bottom flange),
at which the temperature was the lowest, and hence the vertical gradient was zero.
In December 25, vertical surfaces receive considerable amounts of solar energy during the whole day, which leads to the
increase of the temperature of the central thermocouple. However, due to the low conductivity of concrete, the temperature there
still slightly less than the temperature of the top surface thermocouple. As a result, the winter maximum gradient distribution
has three inflection points, whereas it is only one in summer as shown in Figure 6.
ABID ET AL. 9 of 23

FIGURE 6 Distribution of maximum vertical temperature gradients in July and December

3.5 | Vertical daily temperature gradient distributions


The distributions of temperature gradient along the depth of the girder at different time steps from 6:00 to 21:00 are shown in
Figure 7a for July 5 and in Figure 7b for December 25. In July 5, as sunrays start to warm up the girder after sunrise, the
temperature gradient changes from negative with top surface's temperature lower than that of the interiors' to positive with hotter
top surface. At 6:00 in the morning, hence, less than 1 hr after sunrise, the vertical gradient is almost even, neither negative nor
positive, as shown in Figure 7a. Three hours later, at 9:00, the effect of solar radiation is obvious to change the gradient to
positive with a clearly hotter top surface. This effect continues to reach the highest positive gradients during the midday hours

FIGURE 7 Distributions of vertical temperature gradients at specific times (a) in July 5 and (b) in December 25
10 of 23 ABID ET AL.

as shown in the gradient distribution at 12:00. As the sun moves from its highest position in the sky, the temperature of the top
surface starts decreasing showing lower temperature gradients with time as shown in the distributions at 15:00 and 18:00. Dur-
ing the night hours, the temperature of the top surface decreases whereas the interiors of the girder keep the most of the heat that
was conducted during the day. As a result, negative vertical gradients form as shown from the gradient distribution at 21:00.
The bottom surface also shows positive and negative gradients as shown in Figure 7a. This is due to the heating from the
reflected radiations from the ground and the direct convection during the cold hours. The central thermocouple (TS2) showed
the zero gradients at all time steps except at those with high positive temperature gradients (12:00 and 15:00) where the zero
gradients occurred within the bottom flange (TS3).
In December 25 and as shown in Figure 7b, similar changes occurred in the vertical temperature gradients but with some
differences. In December 25, the values of the negative gradients are comparable to the positive gradient values; whereas in July
5, the negative gradients are too small compared to positive gradients. Another difference is that in December 25, the gradients
(positive and negative) at the bottom surface of the girder are not small compared to those at the top surface. On the other hand,
in July 5, the positive gradients at the bottom surface are too small compared to the top surface's gradients, whereas they are
almost equal for negative gradients. The shape of the positive vertical temperature gradient in December 25 is also different
from that of July 5 as discussed in the previous section.

3.6 | Lateral daily temperature gradient distributions


The distributions of the summer's lateral temperature gradients along the top flange shown in Figure 8a are interesting. At early
morning (6:00), it is clear that the northern edge temperature is higher than all other parts, whereas the southern edge is the
coldest. This is because the sun rises from northeast as discussed previously. As the sun moves higher in the sky, the temper-
ature of the south edge also increases composing the distribution shown at 9:00. At 12:00, the north edge becomes shaded,
which reduces its temperature compared to the south edge. As the top surface continues receiving high radiations during the

FIGURE 8 Distributions of lateral temperature gradients along the top flange at specific times (a) in July 5 and (b) in December 25
ABID ET AL. 11 of 23

midday hours, the temperatures of the top surface thermocouples TC6 and TC7 become higher than edge temperatures, whereas
the temperature of the central thermocouple TC3 is lower because of the continuity with the colder concrete of the web. As a
result, the distribution at 15:00 forms. This distribution reverses during the night hours due to the convection cooling at the
exposed surfaces and the long‐wave reradiation as shown in the distributions at 21:00.
In December 25, the distributions of lateral temperature gradients along the top flange are too steady compared to those in
July 5 as shown in Figure 8b. As the sun rises from the southeast and continues moving to the south of the equatorial plane, the
temperature of the southern edge of the top flange is always higher than that of the northern edge during the day. Moreover, the
significant changes of the gradient take place only at edge thermocouples; whereas in contrary to the distributions in July 5, TC5
and TC6 showed no significant variations compared to the central thermocouple TC3.
The lateral distributions across the web shown in Figure 9 are simple and similar to those along the top flange in December
25. It is obvious in Figure 9a that the highest temperature gradients in July 5 occur at the northern surface, whereas it is clear in
Figure 9b that the gradients at the southern surface are higher in December 25.

3.7 | Thermal strains


Figure 10 shows the hourly variation of the strains measured from the two strain gages that were embedded at the central depth
of the top flange. The strains were measured as the change of the strains from an initial strain reading. This reading was the
strain value at the start time of each period, which is at the midnight (00:00) of the first day of each period, hence, July 1
and December 21. The data logger was supplied with a compatible strain analyzer, which uses a spectral interpolation method
to measure the resonant frequency. The recorded frequency at each time step was then converted to microstrains after
subtracting the frequency at the starting time (initial frequency) and multiplying by conversion factors provided by the
manufacturer.

FIGURE 9 Distributions of lateral temperature gradients across the web at specific times (a) in July 5 and (b) in December 25
12 of 23 ABID ET AL.

FIGURE 10 Ten‐day variation of thermal strains: (a) July 1 to 10 and (b) December 21 to 30

The nonlinear temperature distributions shown in Figure 6 cause free nonlinear thermal strains that try to deform the
section nonlinearly. However, the actual thermal strain distribution is linear because the section should keep plane after
deformation. As a result, equilibrating thermal strains form. These thermal strains are compression along the hotter regions
and tension along the colder regions. As shown in Figure 10a,b, the strains increased from zero at the starting time to tensile
strains during the cooling hours where the vertical temperature gradient was negative. Starting from the first shining hours,
the thermal strains changed from tension toward compression as the temperature of the top surface increased. The
compression strains increased during the day hours reaching maximum daily values during the midday hours. After which,
the thermal strains reversed toward zero strain as the temperature of the top flange decreased. This cycle is repeated for
the next 9 days, but with different strain values depending on the temperature gradient distributions of these days compared
to the starting time.
A maximum compression strain of approximately 170 microstrains and a maximum tensile strain of approximately 40
microstrains were recorded during the first 10 days of July. During the last 10 days of December, the recorded maximum
compression and tensile strains were approximately 110 and 40 microstrains, respectively. It is worth to remind that these strains
are change of strains from the initial strain values.

4 | F I N I T E E L E M E N T TH E R M A L M O D E L L I NG

4.1 | Heat transfer in the girder


The heat conduction in the composite girder is controlled by the partial differential equation expressed below[39]:

 2 
∂ T ∂2 T ∂2 T ∂T
k þ þ ¼ ρC p : (1)
∂x2 ∂y2 ∂z2 ∂t
ABID ET AL. 13 of 23

In the above Fourier equation, T refers to the temperature at any point in the girder at any time, t, whereas k is the thermal
conductivity coefficient of the material (concrete and steel) in W/m°C, which is equal in all directions for isotropic materials, ρ is
the density of material in kg/m3, and Cp is the specific heat in J/kg°C.
The boundary loads on the surfaces of the girder can be given as

∂T ∂T ∂T
k nx þ k ny þ k nz þ q ¼ 0: (2)
∂x ∂y ∂z

The terms nx, ny, and nz are the direction cosines of the normal vectors,[39] whereas the term q represents the sum of all ther-
mal loads on the boundaries in W/m2, which includes the convection qc, the total solar radiation qs, and the long‐wave reradi-
ation qre, which are given by

q ¼ qc þ qs þ qre : (3)

The convection between the atmospheric air and the exposed surfaces of the girder is represented in Equation 4 and is
dependent of air temperature Ta, surface temperature Ts, and the convection coefficient hc, which is given in Equation 5 and
is dependent of the wind speed w.

qc ¼ hc ðT s −T a Þ; (4)

where

hc ¼ 5:6 þ 4:0w: (5)

The total solar radiation qs refers to the sum of three components. The direct solar radiation qsr (beam and diffused),
which is represented by Equation 7, the ground reflected radiation qgr, and the surface‐to‐surface radiation among the girder
surfaces qss.

qs ¼ qsr þ qgr þ qss ; (6)

qsr ¼ aI s ; (7)

qgr ¼ aI g ; (8)

where Is is the solar radiation received by any horizontal surface on the ground, whereas a is the absorptivity of the
girder surfaces. Ig is that fraction of the solar radiation reaching and reflected by the ground and other objects around
the girder.
 
1−cos δ
I g ¼ ag I s ; (9)
2

where ag is the reflection coefficient of the ground (albedo), whereas δ is the surface inclination angle.
The term qre describes the cooling of the girder via heat emitting by long‐wave radiation. This term changes nonlinearly
based on the difference between the quartic temperatures of the air and the girder surfaces.

qre ¼ ϵβs T 4s −T 4a (10)

In which, ϵ is the emissivity of girder surfaces and βs is the Stefan–Boltzmann constant, which equals 5.67 × 10−8W/m2K4.

4.2 | The FE modelling


The heat transfer problem discussed in the previous section is simulated for the experimental composite girder shown in
Figure 1 using COMSOL 4.3a multiphysics FE package.[40] In COMSOL, it is required to define the type and order of the ele-
ments used in the meshing process. The volume of the girder was meshed using quadratic tetrahedral elements, whereas the
surface (boundary) elements were meshed using triangular elements.
14 of 23 ABID ET AL.

The size of the elements was chosen to be fine enough to capture accurate temperatures. In the mesh of the girder, the total
number of tetrahedral elements was approximately 21,000 elements, whereas the triangular surface elements were approxi-
mately 6,000. The hemi‐cube method of COMSOL was used to calculate the mutual view factors between the surfaces of
the girder to evaluate the surface‐to‐surface radiation. This method is accurate because it considers the shadowing effect of
the different parts of the girder.
Depending on the literature,[33] the material thermal properties were selected carefully to efficiently simulate the thermal
FE analysis of the composite girder. Table 2 lists the material properties that were used. In addition, the used emissivity
and the absorptivity of concrete were 0.85 and 0.5, respectively. The initial time of the FE thermal analysis of the girder
was selected to be the midnight, at which the temperature of the whole girder was considered equal to air temperature at
that time. This time was selected because the temperature differences between all parts of the girder are small after
several cooling hours. The analysis was started 48 hr before the target day to eliminate the effect of the supposed uniform
temperature at the initial time.

4.3 | Verification of the FE temperatures


To verify the conducted FE thermal analysis, 2 days were selected for comparison purposes; these are July 5 and
December 25. The experimentally recorder air temperature at each time step (each 30 min) along these days and their
previous 2 days were uploaded to the FE model. These records were used to calculate the convection cooling and the
long‐wave radiation according to Equations 4 and 10, respectively. The reflected radiations on the lower‐horizontal and
vertical surfaces of the girder were calculated based on the experimentally recorded solar radiations of the 2 days,
considering a ground albedo of 0.25.
The verification was conducted in terms of comparisons between the experimental and the FE temperatures of each thermo-
couple along the 24 hr of the specified day (July 5 or December 25, 2015). Figure 11 shows the comparisons in July 5 for 6
thermocouples distributed in different locations in the girder. As shown in Figure 11, the FE model could accurately capture
both the temperature values and their variation behavior with time. The maximum absolute recorded error between the exper-
imental and the FE temperatures for all of the 15 thermocouples and along the 24 hr of July 5 was 3.2 °C. For December 25 on
the other hand, and because of the low temperatures, lower temperature errors were recorded. The absolute errors in December
25 for all thermocouples were in general less than 2.0 °C.

5 | S I Z E E F F E C T O N T E M P E R AT U R E A N D T E M P E R AT U R E G R A D I E N T S

As reviewed in the introduction section, the design temperature gradients of the bridge design codes are considered appli-
cable for wide range of section sizes. However, it was disclosed in previous researches[2,3,41] that the size of the section
matters and that the available design gradients may result in underestimated or overestimated stresses for large sections or
special configuration sections. From this review, it can be said that the depth of the girder and the thickness of the
flanges and web should have an influence on the temperature gradient distributions. In this section, a parametric study
is carried out to investigate the effect of the size of the girder using the verified FE model. This parametric study is
subdivided into three studies. In the first, the overall section sizing is studied so that both section depth and member
thicknesses are increased in similar rates keeping the same aspect ratios. This parametric study investigates the combined
effect of the girder depth and the thickness of its flanges and web. In the second parametric study, the effect of the girder
depth is studied using fixed thickness of web and flanges, whereas the third investigates the effect of the thickness of the
web and flanges for a fixed depth section. In all parts of the FE parametric study, the environmental data of July 5, 2015,
were used.

TABLE 2 Concrete and steel thermal properties used in the finite element thermal analysis

Material property Concrete Steel

Thermal conductivity (W/m°C) 1.5 44.5


°
Specific heat (J/kg C) 900 475
3
Density (kg/m ) 2,400 7,850
ABID ET AL. 15 of 23

FIGURE 11 Finite element (FE) versus experimental 24‐hr variation of temperature of specific thermocouples

FIGURE 12 Dimensions of the finite


element sized composite girder sections

TABLE 3 Dimensions of the finite element sized composite girder sections

Dimensions (mm) A B C D E a b c

Experimental 800 300 500 100 100 200 400 8


Double 1,600 600 1,000 200 200 400 800 16
Triple 2,400 900 1,500 300 300 600 1,200 24
16 of 23 ABID ET AL.

FIGURE 13 Size effect on 24‐hr temperature distributions at (a) top; (b) bottom; (c) mid‐height; and (d) north edge of the composite girder
ABID ET AL. 17 of 23

5.1 | Experimental, double, and triple size girders


Based on Figure 12, the dimensions of three FE girders are listed in Table 3. The first girder has the same dimensions of the
experimental girder; whereas in the second and the third girders, all dimensions were doubled and tripled, respectively, keeping
fixed geometrical aspect ratios. As shading effects depend on the aspect ratios of the girder members, they were kept fixed in
this stage to overcome any result confusion due to the different shading effects.
Figure 13 shows comparisons of the 24‐hr temperature variation at selected thermocouple locations. The selected locations
are the center of the top concrete surface of the girder, the center of the bottom concrete surface of the girder, the central height
of girder within the steel section, and the center of the northern concrete edge surface of the top flange, which are shown in
Figure 13a–d, respectively.
As shown in Figure 13a, the temperature differences between the three girders are quite acceptable during the 24 hr at
the top‐surface thermocouple. The maximum temperature difference at this point is 4.7 °C, whereas the minimum difference
is less than 0.2 °C. As shown in Figure 13a, the temperature difference starts increasing after sunset and continues increas-
ing during the night cooling hours to reach maximum around the sunrise. During the sunshine hours, the opposite stands,
where the temperature gap between the three girders decreases during this period. The explanation of this behavior is that
for thicker flanges, higher heat is reserved during the day heating hours because of the higher concrete mass. Thus, longer
night cooling hours are required to decrease the temperature of the surface. As a result, the smaller (experimental size) top
surface cooled faster during the night hours showing lower temperatures than the larger girders, whereas it shows slightly
higher temperatures during the mid‐day hours. Figure 13b shows that similar behavior stands for the bottom surface ther-
mocouple but with slightly different rates. The minimum and maximum temperature differences at this thermocouple are
0.8 and 4.8 °C, respectively.
Figure 13c shows that the size of the girder has larger impact on the temperature of the steel central‐web thermocou-
ple than on those at the top and bottom surfaces. As shown in the figure, the largest girder (triple size) shows the lowest
daily temperature fluctuation, whereas the smallest one (experimental size) shows the highest fluctuation. In contrary to
the top and bottom surfaces, the temperature difference is high during both the cold and the hot hours. The maximum
temperature difference of this thermocouple during the 24 hr is 7.8 °C. This is because the larger girders have thicker
webs, hence smaller girders conduct more heat to the steel section during the day hours and allow to faster reversed heat
conduction during the night hours. As a result, the temperature of the steel section of smaller girders increases faster dur-
ing the day and decreases faster during the night. As shown in Figure 13d, the temperature variation between the three
girders at the top‐flange edge thermocouple is also significant with a temperature difference ranges from 0.9 to 8.6 °C
during the 24 hr.
In Figure 14, the distribution of the vertical temperature gradient is used to evaluate the effect of the size of the girder. The
effects of temperature differences between the three girders discussed in Figure 13a–c are reflected here on the vertical temper-
ature gradient distributions. The maximum temperature gradient occurs during the mid‐day hours, where solar radiation and air
temperature are around their daily highest records. At this time and as shown in Figure 13a, the top surface temperature is
almost the same for all girders (between 48 and 49 °C) with slightly higher temperature for the smallest girder. On the other
hand, the temperatures of the steel web are much lower for larger girders as shown in Figure 13c. The lowest temperatures

FIGURE 14 Size effect on vertical temperature gradient


18 of 23 ABID ET AL.

FIGURE 15 Twenty‐four‐hour temperature distributions for different girder depths with 150 mm thick web and flanges at (a) top; (b) bottom; (c)
mid‐height; and (d) north edge of the composite girder
ABID ET AL. 19 of 23

of the steel web of the experimental, double, and triple size girders at the time of maximum gradient (13:00) are 35, 31.3, and
29.9 °C, respectively. This vertical temperature distribution leads to the increase of the vertical temperature gradient as the size
of the girder increase as shown in Figure 14.

5.2 | Effect of the girder depth


In the previous section, it was shown that the size of the girder affects the temperature variations and temperature gradient
distributions. The effect of the depth of the girder is discussed in this section. This FE parametric study compares between
the same sized beams of the previous section but with fixed web and flanges thickness for all girders, which is 150 mm.
Thus, the same dimensions are shown in Table 3, with different depths of 500, 1,000, and 15,000 mm but with
D = E = 150 mm.
Figure 15 shows that for sections with variable depth and fixed member's thickness of 150 mm, the temperature variations
between the 500, 1,000, and 1,500 mm deep girders are in general minimal and much smaller than those shown in Figure 13.
The maximum temperature differences between the three girders at the concrete top surface, concrete bottom surface, steel web,
and concrete top‐flange's edge are 0.8, 1.3, 1.7, and 2.1 °C, respectively. Similarly, Figure 16 shows that for the girders having
depths of 500, 1,000, and 1,500 mm with constant web and flange thicknesses, the maximum vertical temperature gradients are
quite the same, which are 14.8, 15.4, and 15.0 °C, respectively. This leads to the conclusion that if the thicknesses of the girder's
members are kept constant, increasing the height and width of these members with the same aspect ratios has minor effect on
temperature and gradient distributions.

5.3 | Effect of the web and flanges thickness


In this section, the particular effect of the thickness of the web and flanges is investigated. The effect of the thickness of the
girder members is shown in Figures 17 and 18. In these figures, all girders have the dimensions of the triple size girder shown
in Table 3. However, the thickness of the flanges and the webs (D and E) is variable from 150 to 300 mm with a step of 50 mm.
It is shown in Figure 17a,b that as the thickness increases from 150 to 300 mm, the temperature variations increase. This is more
noticeable in Figure 17c,d at the steel web and the edge of the top flange. The maximum temperature differences at the top sur-
face, the bottom surface, the steel web, and the top‐flange edge are 2.7, 3.0, 5.5, and 6.9 °C, respectively. Comparing these
values with those from Figures 13 and 15, it is obvious that the greatest part of temperature variation can be attributed to the
thickness of the concrete web and flanges.
Comparing the vertical temperature gradients of the four triple size girders with different thicknesses of concrete web and
flanges, it is clear that the maximum gradient increases as the thickness increases. The maximum vertical temperature gradients
of the triple size girders with thicknesses of 150, 200, 250, and 300 mm are 14.9, 16.7, 17.2, and 18.5 °C, respectively, as shown
in Figure 18. This result confirms the conclusion that the thickness of the web and the flanges has a significant impact on tem-
perature distributions and temperature gradients.
As a result of the three parametric studies and within the limits of the studied parameters, it can be concluded that
the size of the girder affects the temperature and temperature gradient in girders subjected to solar radiation and air

FIGURE 16 Vertical temperature gradients for different girder depths with 150 mm thick web and flanges
20 of 23 ABID ET AL.

FIGURE 17 Twenty‐four‐hour temperature distributions for different web and flanges thicknesses at (a) top; (b) bottom; (c) mid‐height: and
(d) north edge of the composite girder
ABID ET AL. 21 of 23

FIGURE 18 Vertical temperature gradients for different thicknesses of web and flanges

temperature. The results also showed that the thickness of the flanges and web has more significant effect than that of
the girder depth.

6 | CONCLUSIONS

As one of the temporal loads that affect the long‐term structural performance, temperature variation was generally one of the
major foci of bridges SHM programs. Aiming to investigate the temperature distributions in concrete‐encased steel girders
and the effect of the girder size on these distributions, an experimental work was carried out and a FE parametric study was
conducted. The vertical and lateral temperature gradient distributions were determined from thermocouples instrumented in dif-
ferent locations in an experimental concrete‐encased steel girder segment. The thermocouple records in addition to strains and
other measurements were continued from the start of July to the end of December. The FE thermal analysis was used to inves-
tigate the effect of the depth of the girder, the thickness of the web and flanges, and the combined effect of them on temperature
and temperature gradient distributions.
The experimental results showed that the variations and the magnitudes of the maximum vertical and lateral temper-
ature gradients are highly dependent on sun movement and its altitude in the sky, and hence on the season of the year
and on the location of the structural element with respect to sun movement. High vertical temperature gradients reaching
15 °C were recorded in July, whereas the recorded lateral temperature gradients were less than 6 °C. In contrary, during
December, lateral temperature gradients were more significant than vertical gradients. The recorded maximum lateral tem-
perature gradient in December was more than 12 °C, whereas the recorded maximum vertical temperature gradient was
only 3.2 °C. The distributions of vertical and lateral temperature gradients along the web, along the top flange, and across
the web were also found significantly different between July and December. The variation of thermal strains with time
from compression to tension is directly related to the distribution of the vertical temperature gradient and its variation
with time.
The FE model of the girder showed good capability to simulate the heat conduction, convection, and radiation in the girder
and with the surrounding environment. For the 15 experimental thermocouples, the maximum absolute error with the FE tem-
peratures and along the 24 hr was 3.2 °C in July 5 and 2.0 °C in December 25.
To study the effect of the girder size and using the verified FE model, the girder size was doubled and tripled compared to
the experimental girder keeping the same geometrical aspect ratios. The comparisons between the three girders showed that
smaller girders suffer higher daily temperature fluctuation. At the central depth of the composite girder within the steel section,
the daily maximum temperature difference between the three girders was 7.8 °C. On the other hand, the maximum vertical tem-
perature gradient was higher for larger girders. The maximum vertical temperature gradients of the experimental, double, and
triple size girders were 14.0, 16.7, and 18.5 °C, respectively.
Another FE parametric study was directed to investigate the particular effect of the girder depth. For girders with depths of
500, 1,000, and 1,500 mm, but with constant thickness of web and flanges of 150 mm, the maximum daily temperature vari-
ation was only 2.1 °C, whereas the maximum difference between the maximum vertical temperature gradients of the three
girders was less than 0.7 °C.
22 of 23 ABID ET AL.

A third parametric study using 1,500 mm deep sections with different web and flanges thickness of 150, 200, 250, and
300 mm showed that the largest share of the temperature and temperature gradient variations can be attributed to the thickness
of the web and flanges of the girder. Thus, the effect of the thickness of the web and flanges on temperature variation and tem-
perature gradient distributions is more significant than the effect of the girder depth.

R E F E RENC E S
[1] F. A. Branco, P. A. Mendes, J. Struct. Eng. 1993, 119(8), 2313.
[2] D. Li, M. A. Maes, W. H. Dilger, Can. J. Civ. Eng. 2004, 31(5), 813.
[3] Z. Song, J. Xiao, L. Shen, Adv. Struct. Eng. 2012, 15(3), 399.
[4] R. A. Imbsen, D. E. Vandershaf, R. A. Schamber, R. V. Nutt, Thermal effects in concrete bridge superstructures, in National cooperative highway
research program report 276, Transportation Research Board, Washington, DC, USA 1985.
[5] M. Elbadry, A. Ghali, ACI J. 1986, 83(6), 1001.
[6] F. Giussani, Eng. Struct. 2009, 31(10), 2392.
[7] W. Dilger, J. C. Beauchamp, M. S. Cheung, A. Ghali, J. Struct. Div. 1981, 107(ST11), 2147.
[8] K.‐Y. Wong, Struct. Control Health Monit. 2004, 11(2), 91.
[9] G. W. William, S. N. Shoukry, M. Y. Riad, Bridge Struct. 2005, 1(2), 103.
[10] Q. Chen, Effects of thermal loads on Texas steel bridges. PhD Thesis, the University of Texas at Austin, Texas, USA, 2008.
[11] S. N. Shoukry, M. Y. Riad, G. W. William, Eng. Struct. 2009, 31(12), 2954.
[12] Y. Xu, B. Chen, C. Ng, K. Wong, W. Chan, Struct. Control Health Monit. 2010, 17(6), 632.
[13] K. Koo, J. Brownjohn, D. List, R. Cole, Struct. Control Health Monit. 2013, 20(4), 609.
[14] Y. Xia, B. Chen, X.‐Q. Zhou, Y.‐L. Xu, Struct. Control Health Monit. 2013, 20(4), 560.
[15] E. Obrien, C. Carey, J. Keenahan, Struct. Control Health Monit. 2015, 22(12), 1396.
[16] J. Gu, M. Gul, X. Wu, Struct. Control Health Monit. 2017, https://doi.org/10.1002/stc.1998.
[17] J.‐H. Lee, I. Kalkan, Adv. Struct. Eng. 2012, 15(3), 447.
[18] N. Kulprapha, P. Warnitchai, Eng. Struct. 2012, 40, 20.
[19] R. Kromanis, P. Kripakaran, Comput. Struct. 2014, 136, 64.
[20] S. R. Abid, N. Tayşi, M. Özakça, Constr. Build. Mater. 2016, 106, 523.
[21] S. R. Abid, N. Tayşi, M. Özakça, Experimental measurements on temperature gradients in concrete box‐girder bridge under environmental ther-
mal loads. In Proceedings of the Istanbul Bridge Conference, Istanbul, Turkey, 2014.
[22] Y. Wang, Y. Zhan, R. Zhao, Adv. Struct. Eng. 2016, 19(7), 1043.
[23] Y. Xia, Y.‐L. Xu, Z.‐L. Wei, H.‐P. Zhu, X.‐Q. Zhao, Eng. Struct. 2011, 33, 146.
[24] H. Liu, Z. Chen, T. Zhou, Appl. Therm. Eng. 2012, 37, 329.
[25] C. Berwanger, J. Struct. Eng. 1983, 109(10), 2325.
[26] J.‐M. Lucas, A. Berred, C. Louis, Struct. Build. 2003, 156(2), 175.
[27] O. Larsson, Struct. Concr. 2009, 10(4), 193.
[28] H. Li, S. Li, J. Ou, H. Li, Struct. Control Health Monit. 2010, 17(5), 495.
[29] O. Larsson, R. Karoumi, Struct. Eng. Int. 2011, 21(1), 74.
[30] B. Chen, Z. Chen, Y. Sun, S. Zhao, Math. Probl. Eng. 2013, 2013(11), 1.
[31] D. Bortoluzzi, S. Casciati, L. Faravelli, J. Vib. Control 2013, 19(15), 2301.
[32] B. Gu, Z. Chen, X. Chen, J. Cent. South Univ. 2014, 21(3), 1227.
[33] N. Tayşi, S. R. Abid, Adv. Struct. Eng. 2015, 18(4), 469.
[34] H. Nandan, M. P. Singh, Eng. Struct. 2014, 81, 480.
[35] AASHTO, American association of state highway and transportation officials, AASHTO LRFD bridge design specifications. Washington DC,
USA 2012.
[36] British Standards Institution, BS 5400: Part2. Steel, concrete and composite bridges, specifications for loads. London, UK, 2006.
[37] European Committee for Standardization, EN 1991‐1‐5:2003. Eurocode 1: actions on structures‐part 1–5: general actions‐thermal actions.
Brussels, Belgium, 2009.
[38] NZ Transport Agency, Bridge manual SP/M/022. Section 3: design loading. Wellington, New Zealand, 2013.
ABID ET AL. 23 of 23

[39] A. Ghali, R. Favre, M. Elbadry, Concrete structures: Stresses and deformation, 3rd ed., E & FN Spon, London 2002.
[40] COMSOL Multiphysics v 4.3, COMSOL multiphysics user's guide. Stockholm, Sweden, 2012.
[41] E. Mirambell, A. Aguado, J. Struct. Eng. 1990, 116, 2388.

How to cite this article: Abid SR, Mussa F, Tayşi N, Özakça M. Experimental and finite element investigation of
temperature distributions in concrete‐encased steel girders. Struct Control Health Monit. 2017;e2042. https://doi.org/
10.1002/stc.2042

You might also like