Vous êtes sur la page 1sur 35

1.

INTEGRAL ABUTMENT BRIDGES

1.1 Overview
In geographical areas with low seasonal temperatures and an abundance of snow and
freezing rain, the use of deicing chemicals to maintain dry pavements throughout the winter had
a significant effect on the durability and integrity of bridges built with deck joints. To help
minimize or eliminate these effects, the integral abutment design was introduced. The term
“Integral Bridges” or “Integral Abutment Bridges” is generally used to refer to continuous
jointless bridges with single and multiple spans and capped-pile stub type abutments. When such
construction is used, it is accepted that the continuity achieved by such construction will subject
superstructures to secondary stresses. These stresses are caused by the response of continuous
superstructures to thermal and moisture changes and gradients, settlements of substructures, post
tensioning, etc., as well as due to restraint provided by abutment foundations and backfill against
the cyclic movement of bridge superstructures [i].
The routine use of integral abutments to tie the bridge superstructure to the foundation
piling began in the United States about 30 years ago. The states of Kansas, Missouri, Ohio and
Tennessee are some of the early users. A recent survey [ii] that has been conducted by
researchers indicates that more than 80 percent of the state highway agencies have developed
design criteria for bridges without expansion devices. As of 1990, 11 states are building
continuous integral bridges at lengths up to 300 feet. Missouri and Tennessee report even longer
lengths. In most respects, integral bridges have performed more effectively, because they remain
in service for longer periods of time with only moderate maintenance and occasional repairs.

1.2 Typical Characteristics of Integral Bridges

Figure 1-1 shows steel stringer bridge life cycle, load events and design issues. The components
and interfaces that are designed empirically or with minor rational calculations are shown with a
question mark. Although there are a lot of unknowns in the design procedure, special attention is
given to abutments, girders, and movement systems by the states.
The main two attributes that have been given special attention in integral bridges are the
abutment details and the allowable bridge lengths. Most of the states that use integral abutments
have developed specific guidelines for both of these attributes. The basis of these guidelines,
however, is largely empirical. In Figure 1-2, four abutment details for short span steel stringer
bridges from different states are given. Since structural details from early successful designs are
adapted by other bridge engineers, there can be seen many similarities as well as differences
among the details.
Although there is a wide variety of length limitations used by many different states due to
a lack of well designed research on this area, the following limitations are commonly used: for
steel superstructure, 200 to 300 ft., for concrete 300 to 400 ft., and for prestressed concrete 300
to 450 ft. The regulations in some states, such as Tennessee, let designers use longer lengths. It
would be desirable to develop these details and limitations according to results of some rational
analysis or tests conducted to find the anticipated movement of bridges and the abutments.
Now that the popularity of these type of bridges has grown, it has become evident that
these bridges have many more attributes and fewer limitations than their jointed counterparts
[iii]. To list a few of these attributes:
• Jointless construction reduce the destructive effects of leaking deck joints
• Reduced cost in future modification and replacement
• Resistance to pavement pressure
• Simple design
• Rapid construction
• Broad-span ratios
• Earthquake Resistance
• Improvement in Live Load Distribution
Some of the “perceived” limitations listed are as follows:
• High Abutment-Pile Stresses
• High passive pressure
• Limited applications
• Buoyancy
To minimize the high abutment-pile stresses and passive pressure developed in abutments,
design engineers have used a number of controls, devices and procedures [i] including but not
limited to the following:
• Limitations on bridge length, structure skew and vertical penetration of abutments into
embankments;
• Use of granular or uncompacted backfill
• Provisions for approach slabs to prevent vehicular compaction of backfill
• Use of semi-integral abutments to reduce passive pressure
• Use of prebored holes filled with fine granular material for piles
• Provisions for an abutment hinge to control pile flexure

1.2.1 Future Trends


There is an increasing tendency in the Department of Transportation towards the use of
integral abutments. In the near future, it is expected that integral abutments will be the standard
for new designs of short to medium span bridges; also, several other techniques will be used.
A recent document, prepared for the American Iron and Steel Institute by J. Muller
International [iv] on innovative designs for short and medium span bridges, envisions integral
abutments as the future standard. In addition, integral piers in which intermediate bearings are
eliminated are stated as excellent seismic details. In this detail, the pier shaft and cap are tied
directly into the superstructure.
For existing bridges, integral conversions will be routinely used for retrofitting and
strengthening purposes. Although it has been rarely done, it has been reported that conversion of
non-integral to integral or semi-integral abutments for both single and multiple span bridges has
begun [i]. Figure 1-4 gives design details for recent integral conversions by ODOT.
Reconstruction of these abutments was made necessary by the damage induced by pavement
growth and pressure, and by deicing chemical deterioration.
1.3 Long-Term Environmental Response Monitoring

1.3.1 Gage Characteristics Relevant to Long Term Measurements


The gages used for long-term measurements were VSM-4000 vibrating wire type strain
gages. The laboratory testing previously conducted on these gages revealed that they do not
require any correction for temperature. In other words, the gages are blind to strains due to
temperature in which no stress are created. So after the readings are acquired, special interest
should be given to understand the movement mechanics of the bridge and the stresses created by
these movements. The concept is explained in Figure 1-5.
When a ∆T Temperature differential is applied uniformly to a steel beam of unit length,
with a thermal expansion coefficient of αt, the elongation of the beam will be εt, which is ∆T* αt.
Since the gage is made from the same material, and there is no stress generated by this
elongation, the reading that one would get by the gage will be zero.
In the third row of the first column of the same figure, the other extreme is shown. Let’s
assume that the beam is completely fixed at the ends. With the application of ∆T temperature
differential, the actual elongation of the beam is zero. But the stress accumulated in the beam is -
εt*E. Although there is no strain in the cross-section, the gage will indicate a strain of -εt, which
is completely stress related.
In the forth row, the same beam is shown with an intermediate situation in which one of
the supports is replaced by linear springs. In this case, the actual strain in the cross-section is ε1,
but the gage reading will indicate a strain of -( εt −ε1), which again will result in the stress
accumulated in the cross-section when multiplied by the elasticity modulus.
It is possible to change even this simple model to better represent the actual situation. In
the second column of Figure 1-5, a temperature gradient is added to the structure. The top of the
structure is warmer than the bottom, which is the actual case most of the time. Naturally the
beam curves up due to this temperature gradient. Also, if we add springs of different stiffnesses
and another layer with different temperature characteristics to represent the concrete, we get the
simple model that is shown at the bottom section, which is most likely to represent the actual
mechanics of the single beam at the structure.
The concept of the strain gages being blind to strains due to temperature in which no
stress are created is explained further in Figure 1-6. In this figure, actual strain readings from the
abutment section are used. Three 24-hour periods are selected for demonstration. The top graph
shows the temperature distribution at this cross-section. As mentioned, the top of the section is
significantly warmer than the bottom. This causes some additional stress that will be discussed in
the following sections.
The bottom graph shows the strains that can be extracted from a single strain gage. The
actual reading obtained from the gage is indicated as the stress-induced strain. The non-
restrained scenario is determined with the help of the thermistor that is built into the gage. The
temperature change is multiplied by the expansion coefficient for steel; this assumes that the
girder is not subject to any restraints. The resultant shows the actual behavior of the girder and is
obtained by adding the two previous curves.
For simplicity from this point forward in the report, unless stated otherwise, all values
given and graphed as strains are stress inducing. When finding the mechanics of the bridge and
abutments, total strain (the resultant strain from temperature and stress inducing strains) is used.

1.3.2 General Behavior


A long-term environmental test is conducted to monitor the temperature and
environmental effects upon the bridge. Temperature and related strain readings, at locations
shown in Figure 1-7, are monitored continuously to capture the temperature and stress
distributions during ambient temperature changes. Monitoring has been successful for two years,
beginning in November, 1994.
Figure 1-8 shows the general trend and the maximum changes encountered during two
years at two locations. The two locations graphed are the abutment location on Girder 0, and the
midspan on Girder 2. The maximum change in the ambient temperature is 111.4 F (-12.3 - 99.1
F). The maximum strain fluctuation during these two years was 359 µε, which corresponds to
10.4 ksi.

The characteristic that appears immediately is the nature of the behavior. As it can be
seen in Figure 1-8 and Figure 1-9, the behavior is the combination of two different cycles with
different frequencies. The first one is the seasonal changes or the yearly cycle that is denoted as
monthly averaged values in Figure 1-9. This cycles’ frequency is roughly a year. The second is
the daily fluctuations in strain due to changes in daily temperatures. The frequency of this cycle
is roughly a day. We will examine these cycles separately since they exhibit different behaviors.

1.3.3 Yearly Responses due to Seasonal Changes


One trend immediately noticed examining the two-year strain data is the decrease of
seasonal fluctuations when moving away from the abutments. As it can be seen in Figure 1-10,
there is significant seasonal fluctuation at the abutment and quarter span, but less fluctuation at
the pier and midspan. On the other hand, the daily fluctuations due to temperature, which will be
explored in proceeding sections, are clearly evident at all instrumented locations.
In order to get a better appreciation of seasonal effects, the data is averaged monthly and
each month is represented by a single data point. The results are presented in Figure 1-11. The
top graph indicates the temperature difference between the bridge and the ambient. Also, the
temperature gradient between the top and the bottom of the cross-section can be seen and
appears more significantly during summer months.
In the second graph, stress accumulations and the seasonal effects that were mentioned in
the previous paragraph are seen. The abutment section accumulates the most stress. Amount of
stress and the seasonal cycle magnitude reduces as one approaches the midspan. Table 1-1
summarizes the stress differentials through two years of monitoring. The values reported are:
• Column A Maximum stress differential (max. stress-min. stress)
• Column B Maximum stress differential of curve fit (yearly cycle)
• Column C Daily cycles (3 x Std. Deviation of all stress values within the month, 99%
confidence interval).

A similar study is done for the temperature readings and it is presented in Table 1-2.
1.3.3.1 Abutment Kinematics
The bottom graph in Figure 1-11 represents the total strain at girder 2. As mentioned
before, total strain is the resultant of temperature strains and stress inducing strains. By
integrating the total strain values along the bridge, it is possible to get a rough estimate of the
total elongation of the bridge structure. From the 2-year total strain readings, 1.2” maximum
elongation was calculated. On average, 0.72” total elongation in bridge length was calculated
between January 96 and August 96. Assuming the total elongation is divided between the two
abutments, each abutment moves 0.36”.
In an attempt to understand the behavior of the structure with respect to temperature, the
strain values for girder two is plotted against temperature (Figure 1-12). Again, the relationship
is most apparent and distinct at abutment. Also, the different slopes (µε/T) of the top and bottom
gages indicate a bending action especially on abutment and the pier sections. As one approaches
the midspan, the slopes become parallel, indicating more of an axial deformation.
Although it is possible to find a linear relationship between temperature and strain for
two years, it is conceived that the relationship is an overall behavior and it is not representing the
daily or the short-term behavior of the bridge with temperature changes. This assumption is
further explained in Figure 1-13. In this figure, one week of data from the indicated months are
taken and plotted as microstrain vs. temperature for abutment sections on Girder 0 and Girder 2.
The response greatly varies within the year, significantly increasing in the warmer months at
bottom flanges.
To identify this behavior of the bridge, the change in strain per degree Fahrenheit change
in daily temperature cycles are shown in Figure 1-14. The graph should not be confused with
strain responses, which are shown in Figure 1-11. In Figure 1-14, the parameter under
investigation is the slope (∆µε / ∆ temp). The strain response per degree Fahrenheit change
within daily temperature cycles is not constant throughout those two years. While all the top
gages are exhibiting the same behavior, the bottom flange gages show varying behavior. As the
temperature increases, the tension response of the top gages increases. For bottom flange gages,
as the ambient temperature increases, the abutment sections accumulate more compression per
degree Fahrenheit change. The quarter span remains almost constant and the pier and the
midspan show increasing tension response within the daily temperature cycles.
Three reasons are hypothesized for the behavior shown in Figures 1-11 through 1-14 and
discussed in the paragraphs above. They are:
• The detail that connects the encasement beam to the footing,
• The passive pressure generated by the abutment backfill, and
• The temperature gradient between the top and the bottom of the section.
The detail connecting the encasement beam to the footing can be seen in Figure 1-15. The
lateral movement of the encasement beam is restrained by the #5 rebar that is connecting the two
pieces. Therefore, high stress concentrations at the abutment section near the bottom flange are
seen. The reasoning behind this detail is that creating a rotational degree of freedom will reduce
the stresses at the abutment piles. The design idealizations and the hypothesized behavior can be
seen in Figure 1-15. In the undeformed detail, the connection between the encasement beam, the
footing and the approach slab is shown. The design assumption is that the footing will take all
the deformation under temperature loading. From the readings, it is hypothesized that the
encasement beam as well as the footing are under displacement and rotation. Without extensive
instrumentation and modeling, it is difficult to visualize the actual behavior of the abutment.
Hence, for better design details, additional research should be carried out.

1.3.3.2 Passive Pressure


In many publications[i,ii,v,vi,vii], the passive pressure developed in abutment backfill by
an expanding integral bridge is identified as a problem to be solved. Design engineers have used
a number of controls to minimize the effects. These controls include, but not limited to:
• Limiting the bridge length, skew, and vertical penetration of abutments into embankments
• Using granular or uncompacted backfill
• Using approach slabs to prevent vehicular compaction of backfill
• Using semi-integral abutments (Figure 1-15)
The responses seen in Figures 1-11 to 1-14 can be a result of a change in the boundary
conditions; especially, pressure generated by the abutment backfill will change with time,
temperature and moisture. Although granular backfill is indeed used to minimize the passive
pressure for this bridge, the compression (Figure 1-11) and the compression rate per degree
Fahrenheit (Figure 1-14) is increasing in summer months at the abutment sections.
Elgaaly et. al.[vi], reports that the 6-month behavior of the Forks Bridge in the state of
Maine shows similar results. In this research, the soil pressure behind the abutments was
measured. It was found that temperature variation affects soil pressure because of the resulting
bridge deck expansion and contraction.
Burke [iii], recommends extensive research on passive pressure, to describe both the
relationship between the amount of soil compression and the generation of passive pressure and
the effects of alternating cycles of soil compression and expansion.

1.3.3.3 Bridge Deformation Under Temperature


The discussion in the preceding sections also brings into question the curvature change of
the different sections of the structure with temperature change. This can be used to have a better
estimation of the deformation kinematics of the structure under ambient temperature loading.
In Figure 1-16, curvature values for monthly averaged data for two years are plotted.
Since we do not have an absolute zero value when we can say that the bridge is perfectly
straight, an arbitrary zero value was chosen as the start of sampling (Nov. 94). Because of that
reason, the graphs should be an indication of trend instead of absolute deformations. The top
graph shows the curvature change vs. temperature along Girder 2. The graph shows a negative
change in the curvature with increasing temperature at the sections of abutment and 1st span. The
rate of change is higher at the abutment. Over the pier, an opposite change in curvature is seen.
Midspan section shows almost no bending, meaning the deformation is in the form of axial
elongation. The second graph shows the curvature change at the abutment section. All three
girders show similar trend: decreasing curvature with increasing temperature. The magnitude of
the change decreases as one goes from the acute angle to the obtuse angle at the skew.
Figure 1-17(a) shows the hypothesized deformation under increasing temperature from
the data collected from strain gages. In developing the figure, the curvature values shown in
Figure 1-16 are used. Again, one of the reasons causing this type of deformation kinematics is
due to the restrained lateral movement of the encasement beam by the #5 rebar that is connecting
the encasement beam and the footing (Figure 1-15). The writers believe, with minor
modifications to the existing detail, it is possible to change the deformation kinematics near the
abutment section. Specifically, using the semi-integral abutment detail shown in Figure 1-15 will
reduce the amount of lateral restraint, causing less rotation, leading to less stress in the deck.
A finite element simulation is performed to examine the creditability of the deformation
kinematics discussed above. A single girder of the bridge is modeled in 3D by using shell
elements and frame elements for concrete and steel members respectively. The soil and the
bearings are modeled by linear and rotational springs. Similar stiffnesses to the calibrated FEM
model are used. 50oF differential temperature is applied to girders and deck as loading. The
deflections obtained from two analyses are shown in Figure 1-17(b). In the top, stiffer linear
springs are used for bearings. The deflection profile obtained from this model reasonably
simulates the hypothesized deflections from the gage readings. Since the model is very sensitive
to the stiffnesses of the supports, several analyses were performed to see the effect of these
parameters. The bottom figure shows the deflection profile when the spring stiffnesses are
reduced. Curvatures and the deflections especially at the midspan are considerably increased in
this model. Since this simple model is not representing the whole bridge, the intention behind
these analyses is not to have a calibrated model; however, the models show that the data gathered
from the gages are credible and no blatant instrument or analysis error exists.
A better understanding of these kinds of bridges can be gathered by application of tilt and
displacement measurement techniques. Although short-term measurement of tilt and
displacements were successful, the researchers did not accomplish reliable long-term
measurement of these parameters. Also, simulation of the effects of different boundary
conditions under temperature loading on a full FEM model can improve the way we approach to
the design of these bridges. It is desirable that additional research should be conducted to provide
a more detailed behavior analysis and to verify the hypothesis stated in this section.

1.3.4 Daily Responses due to Temperature


In order to have a better appreciation of the daily mechanisms seen at the bridge due to
temperature, certain weeks are highlighted and graphed separately. The months of January and
August are selected in which minimum and maximum ambient temperatures occur respectively.
Figure 1-18 shows the daily temperature and strain changes at Girder 2. It is evident from the top
graph that bridge temperatures are higher than the ambient temperature. That is a trend that can
be seen in summer months (Figure 1-19) as well as winter months.
As stated before, the strain change per degree Fahrenheit temperature change in winter is
not as large as the summer months. The top flange gages in both seasons have similar
characteristics. All sections move in similar manner without any phase lag between them. The
magnitudes of the top flange strain gages at different sections are also similar, in both January
and August.
The difference appears when one looks closely to the bottom flange strain readings. In
January, like the top flange gages, the bottom flange gages at different sections follow the same
trend. The gages indicate tension with dropping temperatures. In August, as can be seen in
Figure 1-19 bottom graph, phase and magnitude differences between the gages are apparent.
Abutment and quarterspan sections follow the same trend (i.e., peaking strains at the lowest
temperature). This means that the sections go into compression as temperature increases. But the
pier bottom gage exhibits an opposite behavior, increasing tension with increasing temperatures.
Midspan section follows the abutment behavior with a slight phase difference but with a much
smaller magnitude.
To compare the daily bridge deformations under temperature with the yearly behavior,
daily curvature changes in summer and winter are plotted in Figures 1-20 and 1-21. The same
weeks in which the strain values were graphed are selected. As it can be seen from Figure 1-20,
the curvature changes are very small with the temperature changes. The curvature change for the
same differential temperature in August is more discernible. More bending or bowing action is
seen during summer months than winter months. The curvatures found from daily responses in
August, also supports the hypothesis stated previously about the deformation of the bridge under
temperature.
The maximum temperature change that occurred within a single week is plotted in Figure
1-22. The ambient temperature differential during that week was 76.58oF. That temperature
differential caused a rapid strain accumulation especially at the abutment section. The maximum
strain differential encountered was 287 µε (8.42 ksi) at Girder 0 bottom flange location.
1.4 Design Codes and Recommendations

1.4.1 Thermal Forces


Although AASHTO specifications and several other state design guides contain adequate
information about the flexural and shear design of highway composite bridges subject to dead or
live loads, very little information is given on integral abutment design, thermal loading, design
for thermal loading, and the factors that needed to be included in the bridge design to account for
deterioration with time[viii].
Temperature rise or drop during a day or throughout a year induces longitudinal and
transverse stresses that are often overlooked during the design. The nonuniform temperature
gradient through the deck depth leads to stresses that can reach to high levels as seen in previous
sections. AASHTO specifications do not consider temperature variations through girder depth.
They only specify a rise or drop of mean temperature of the bridge:
Provision shall be made for stresses or movements resulting from variations in
temperature. The rise and the fall in temperature shall be fixed for the locality in which
the structure is to be constructed and shall be computed from an assumed temperature at
the time of erection. the range of temperature shall be as follows:
Metal structures: Moderate climate, from 0 to 120o F. Cold climate, from -30 to 120o F.
[AASHTO Standard Specifications for Highway Bridges Section 3.16]

As mentioned in earlier sections a deck subjected to temperature distribution through


depth experiences three effects: a) expansion and contraction, b) a curvature c) stresses that are
developed because of the restraints and to compensate for the Navier-Bernoulli assumption. New
design guidelines should be developed for deck and girders to accommodate for these effects.
Although the temperature ranges in the draft LRFD Bridge Design Specifications are kept
the same with AASHTO specifications, the vertical temperature gradient in concrete and steel
with concrete decks are taken into account (Section 3.12.3). Also, where temperature gradient is
considered, internal stresses and structure deformations due to both positive and negative
temperature gradients are determined considering axial extension, flexural deformation and
internal stresses (Section 4.6.6).
1.4.2 Integral Abutment Design
Although today more than half of the state highway agencies use integral abutments, the
design practices vary greatly. Little experimental and theoretical work has been conducted to
establish limits or to develop design procedures for integral abutment bridges. This partly
explains the wide variation in the design criteria used by the various state highway agencies.
Surveys show that most of the design practices depend on experience gained on former
designs [ii]. Most of the states that use integral abutments have developed specific guidelines
abutment details (Figure 1-2). No state, except North Dakota, has addressed the effect of passive
soil pressure that acts behind the abutments during the bridge expansion; and almost no state has
addressed the stress concentrations for the girders at abutment sections.
The writer believes that, with minor changes in abutment details, most of the problems
attributed to the integral bridges and discussed previously can be solved. Some of these
modifications are already available in several bridge designs as semi-integral detail as seen in
Figure 1-15 and discussed in following paragraph. It is possible that this type of detail will
reduce the rotation transferred to the abutment and reduce the stress concentrations at the
abutment sections.

1.4.2.1 Semi-Integral Details


As mentioned previously, semi-integral details may solve many shortcomings of integral
abutment details while still utilizing many positive aspects. The basic feature of the semi-integral
abutment detail different from the integral abutment detail is that the superstructure moves
longitudinally on elastomeric bearings almost independent of rigid abutment foundations. The
advantages to this type of detail are [v]:
• Abutment members, including piling, can be designed to operate well within the usual
allowable stress limits;
• Superstructure end areas are reduced resulting in less passive pressure and pressures that are
less eccentric with respect to the neutral axis of the superstructure;
• Abutment and end diaphragm configurations that are simple to design, simple to reinforce,
and relatively simple to construct.
As a result, the semi-integral bridge concept should extend the application range of
bridges with jointless decks to most applications when properly designed and constructed.
Table 1-1 Yearly and Daily Stress Differentials
300

250

200 C

150
Microstrain

100

50
B A
0

-50

-100

-150
S ep-94 Jan-95 A pr-95 Jul-95 O ct-95 Feb-96 M ay-96 A ug-96 D ec-96
Tim e

Max. Stress Max. Monthly Max. Diff.


Location Differential Averaged Stress Daily Cycles
(ksi) Diff. (ksi) (3x St. Dev)

A B C
Abutment Girder 0 Bottom 10.70 4.64 5.85
Top 3.74 1.93 1.63
Girder 2 Bottom 6.13 3.11 3.10
Top 4.01 1.92 1.68
Girder 5 Bottom 2.93 1.20 3.25
Top 6.04 2.92 1.82
Q.Span Girder 2 Bottom 5.04 1.98 3.04
Top 3.86 1.11 1.84
Pier Girder 2 Bottom 4.12 1.02 1.99
Top 4.43 1.94 1.93
Midspan Girder 2 Bottom 4.51 1.10 2.31
Top 3.46 1.41 1.55

Table 1-2 Yearly and Daily Temperature Differentials


Max. Temp. Max. Monthly Max. Diff.
Location Differential Averaged Temp. Daily Cycles
(F) Diff. (F) (3x St. Dev)
A B C
Abutment Girder 0 Bottom 96.2 54.2 46.46
Top 96.0 57.3 46.15
Girder 2 Bottom 92.2 53.8 46.26
Top 96.3 56.3 45.98
Girder 5 Bottom 93.1 53.9 47.04
Top 95.2 56.8 46.50
Q.Span Girder 2 Bottom 96.2 54.0 49.65
Top 99.2 56.2 49.60
Pier Girder 2 Bottom 94.7 53.1 49.92
Top 98.9 55.5 49.56
Midspan Girder 2 Bottom 96.3 52.5 50.00
Top 98.0 54.7 49.77
Ambient Temperature 111.4 55.5 54.02
LIFE-CYCLE LOAD DESIGN
EFFECTS SYSTEMS
CONSTRUCTION HYDROLOGY
Fabrication
DEAD LOADS SOIL
Erection

Construction EMBANKMENT
? ?
SERVICE LOADS PAVEMENT
INTRINSIC
?
Climate & FORCES,
Environment DEFORMATIONS ABUTMENTS ?
& CHANGES IN
Seasonal: THESE FOUNDATIONS ?
Soil & ?
Environmental
Changes ? PIERS/CAPS
?
Changes in
BEARINGS
Ambient condition
?
Traffic GIRDERS ?
LIVE LOADS
Hydrological ?
? X-FRAME ?
DAMAGE
Deterioration DECK ?
Accident
CAPACITY
MOVEMENT ?
Flood REDUCTION
Earthquake DRAINAGE

Figure 1-1 Steel-Stringer Bridge Life-Cycle, Load Events and Design Issues
Integral Abutments: Illinois Integral Abutments: North Dakota

Integral Abutments: Ohio Integral Abutments Tennessee

Figure 1-2 Integral Abutment Details


Defect Deterioration Damage Performance
Deficient Loss of Chemical Deck
Reduced
Interface Bond at Cracking &
Serviceability
Detail Interface: Highly Delamination
Increased Girder Reduced Service
? Stresses Life
Accentuated
Fatigue & Increase at
High Response Corrosion Lifecycle Cost
Excessive Amplification
Flexibility

Large Skew Vertical-Lateral


Coupling and Bearing Drift; Increased
High Response Unseating Vulnerability
Amplifications

Figure 1-3 Bridge Failure Mode Analysis Example (Steel Stringer Bridges)

Figure 1-4 Integral Abutment Conversions


gage reading: −(εt(top)−ε1)
L=1 unit ∆T=T1 total strain: ε1
Ltop=1+ε1
∆T=0 T1>T2
Lbot=1+ε2
∆T=T2
gage reading: −(εt(bot)−ε2)
gage reading: 0 εt total strain: ε2
∆T=T ∆T=T1
total strain: εt
k1 k2
T1>T2

gage reading: −εt ∆T=T2


∆T=T
total strain: 0
∆T=T1

T1>T2
gage reading: −(εt−ε1)
∆T=T total strain: ε1 ∆T=T2

Figure 1-5 Strain Reading at a Beam due to Temperature Changes


95 Temperature Distribution

90

85

80
Temp (F)

75

70

65
Bottom Flange
60 Top Flange HAM-42-0992
WESTBOUND
Ambient Temp
55
8/2 0:00 8/2 12:00 8/3 0:00
Date
00
Comparison of Temp. Strains and Stress Related Strains
100
Non Restrained
80 Stress Induced
Resultant
60

40
Microstrain

20

-20

-40

-60

-80

-100
8/2/96 8/2/96 8/3/96 8/3/96 8/4/96 8/4/96 8/5/96
Date

Figure 1-6 Comparison of Different Strains in the Section


PLAN VIEW
Brg. Abut. 2 Pier 2 Pier 1
girder 0

HAM-42-0992
Westbound
girder 2

girder 5

VIBRATING WIRE GAGE CLUSTER

Figure 1-7 Instrument Locations for Long Term Monitoring


100
Ambient Temp. DEGF
80
Temperature (F)

60

40

20 ∆T=111.4 F

MENT
0

AN
PIER

MIDSP
AB UT
250
∆ε =359 µε
200 ∆σ =10.4 ksi
150
Microstrain

100

50

-50

-100

Sep-94 Jan-95 Apr-95 Jul-95 Oct-95 Feb-96 May-96 Aug-96 Dec-96


Date

Figure 1-8 Two Year Continuous Monitoring Results (Nov 94-Nov 96)

300 60

40
24 Hours
20
250 0
Microstrain

-20

-40
200 -60

-80

-100
150 8/2 0:00 8/2 12:00 8/3 0:00
Microstrain

100

50

monthly average
-50
daily cycles
-100 daily cycle band
(3x std. dev.)
-150
Sep-94 Jan-95 Apr-95 Jul-95 Oct-95 Feb-96 May-96 Aug-96 Dec-96
Time

Figure 1-9 Bridge Behavior Modes during Two-Year Monitoring (Nov 94-Nov 96)
100

Temperature (F) 80

60

40

20
ATEMP DEGF
0 HAM-42-0992
WESTBOUND
-20 02 12 22 32 Feb-96 May-96 Aug-96 Dec-96
150
100 VW02S1
Microstrain

50

-50

150
100 VW12S1
Microstrain

50

-50

150
VW22S1
100

50
Microstrain

-50
-100
150
VW32S1
100
Microstrain

50

-50

-100
Sep-94 Jan-95 Apr-95 Jul-95 Oct-95 Feb-96 May-96 Aug-96 Dec-96
Time

Figure 1-10 Two Year Strain Data for Girder 2


90

80 Temperature

70
Temperature (F)

60

50

40

30
amb temp
20 TH22 Bot
TH22 Top
10
Sep-94 Jan-95 Apr-95 Jul-95 Oct-95 Feb-96 May-96 Aug-96 Dec-96
Date
40
Stress Inducing Strain at Girder 2

20

0
Microstrain

Sep-94 Jan-95 Apr-95 Jul-95 Oct-95 Feb-96 May-96 Aug-96 Dec-96


-20

-40 VW02 bot


VW02 top
VW12 bot
-60 VW12 top
VW22 bot
VW22 top
-80 VW32 bot
VW32 top
HAM-42-0992
WESTBOUND
-100 02 12 22 32
Date
250
Total Strain at Girder 2
200
VW02bot
VW02top
150 VW12bot
VW12top
Microstrain

100 VW22bot
VW22top
VW32bot
50 VW32top

0
Sep-94 Jan-95 Apr-95 Jul-95 Oct-95 Feb-96 May-96 Aug-96 Dec-96
-50

-100

-150 Date

Figure 1-11 Monthly Averaged Values at Girder 2


150 160

Girder 2
bottom 140
Girder 2
100 @ Abutment 120 bottom
@ Q.Span
y = -2.0609x + 142.85 100

50 80
Microstrain

Microstrain
y = -1.2975x + 120.31
60
top
0 40
top
20

-50 y = -1.0112x + 51.346 0

y = -0.6093x + 52.771
-20

-100 -40
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Temp (F) Temp (F)
140 140

120 Girder 2 120 Girder 2


100 @ Pier 100
@ Midspan
bottom y = -0.3441x + 55.696 80
80
bottomy = -0.5889x + 68.608
60 60
Microstrain
Microstrain

40 40

20 20

0 0

-20 -20
y = -1.1313x + 83.386 y = -0.7006x + 50.649

-40 top -40 top


-60 -60
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Temp (F) Temp (F)

Figure 1-12 Strain-Temperature Relationship for Two Years


Temperature vs Strain @ Abutment Girder 0
250
y = -4.23x + 271.62 (Jan 95)
200

y = -4.78x + 308.78 (April 95)


150
y = -5.65x + 429.37(Oct 95)

100 y = -7.60x + 627.26(Aug 95)


Microstrain

50

y = -1.08x + 58.49(Jan 95)


0
0 10 20 30 40 50 60 70 80 90 100
-50 y = -1.36x + 79.82 (April 95)
y = -1.12x + 58.88 (Oct 95)
Abut. G.0. Bottom Fl.
Abut. G.0. Top Fl. y = -1.34x + 83.51 (Aug 95)
-100
HAM-42-0992
WESTBOUND
-150 02
Temperature (F)
00

Temperature vs Strain @ Abutment Girder 2


150
y = -2.06x + 143.05(Jan 95)

100 y = -2.71x + 171.95 (April 95)

y = -3.77x + 274.24 (Oct 95)

50
Microstrain

y = -4.69x + 364.18 (Aug 95)

y = -1.21x + 61.42(Jan 95)


0
0 10 20 30 40 50 60 70 80 90 100
y = -1.33x + 76.45(April 95)
y = -0.88x + 37.34 (Oct 95)
-50
Abut. G.2 Bottom Fl.
Abut. G.2 Top Fl. y = -1.52x + 97.20 (Aug 95)

-100
Temperature (F)

Figure 1-13 Single Week Strain vs. Temperature in Different Months


Monthly Average Temperature over Two Years
90

80
brid. temp
Average Monthly Temp

amb. temp
70

60

50

40

30

20

10
Sep-94 Jan-95 Apr-95 Jul-95 Oct-95 Feb-96 May-96 Aug-96 Dec-96
Month
Change of Strain with ∆T Change in Temperature @ Top Flange
1

0
Slope (∆ µε/∆ Temp)

-1

-2
Abut 0
Abut 2
-3 HAM-42-0992
Q.Span 2 WESTBOUND
Pier 2 02 12 22 32
-4 M.Span 2

-5 00

-6
C h a n g e o f S tr e s s w ith ∆ T C h a n g e in T e m p e r a tu r e @ B o tto m F la n g e
1

-1
Slope (∆ µε/∆ temp)

-2

-3

-4
Abut 0
Abut 2
-5 Q .S p a n 2
P ie r 2
M .S p a n 2
-6
S e p -9 4 J a n -9 5 A p r-9 5 J u l-9 5 O c t-9 5 F e b -9 6 M a y-9 6 A u g -9 6 D e c -9 6
M o n th

Figure 1-14 Change of Strain with DT Change in Temperature


UNDEFORMED DESIGN ASSUMPTION
Approach Slab

Backfill
Encasement
Beam
Exp. Joint Filler

Footing
θ1

HYPOTHESIZED MOVEMENT SEMI INTEGRAL MOVEMENT

θ1

θ1 < θ2 (??)

θ2

Figure 1-15 Integral Abutment Detail Under Temperature Loading


Curvature Change vs. Temperature @ Girder 2 over Two Years
2.00
VW02 ABUT.
VW12 Q.SPAN
1.50 VW22 PIER
VW32 M.SPAN
Curvature (E-6)

1.00
Sign Convention

+
0.50
-
0.00

-0.50
10 20 30 40 50 60 70 80
Temperature (F)
Curvature Change vs. Temperature@ Abutment over Two Years
4.00

3.50
VW00 G.0
Curvature (E-6)

3.00 VW02 G.2

2.50 VW05 G.5

2.00

1.50

1.00

0.50

0.00

-0.50
10 20 30 40 50 60 70 80
Temperature (F)

Figure 1-16 Curvature Change vs. Temperature over Two Years


UNDEFORMED

∆T = +

Figure 1-17(a) Hypothesized Deformation Under Increasing Temperature (not to scale)

55' Z
Concrete: 8.5" ∆T=50 F Y
E=4050 ksi
78' X
α=5.5 E-6 W36x150
55'
W36x170
αsteel=6.5 E-6
W36x150
DEFLECTION
Z .2250E-01

MOMENT
SPRINGS
Z -.5464E-02
Lin: 5E4 K/in
Rot: 1E4 K-in/rad

DEFLECTION
Z .3056E-01
SPRINGS
Lin: 1E4 K/in MOMENT
Rot: 1E4 K-in/rad
Z -.1835E-01

Figure 1-17(b) Simulation of Hypothesized Deformation Under Increasing Temperature


Temperature Distribution @ Girder 2
50

Bridge Temp
40
Temperature (F)

30

20

10 Ambient Temp

0
1/21/96 1/22/96 1/23/96 1/24/96 1/25/96 1/26/96 1/27/96 1/28/96
Date
D a i ly S t r a i n C h a n g e s @ G i r d e r 2 T o p F la n g e
120
ABUT 2

100 QSPAN 2
P IE R 2

80 M ID S P A N 2
Microstrain

60

40

20

-2 0
D a ily S tr a in C h a n g e s @ G ir d e r 2 B o tto m F la n g e
120

100

80
Microstrain

60

40

20
ABUT 2
0 QSPAN 2
P IE R 2
M ID S P A N 2
-2 0
1 /2 1 /9 6 1 /2 2 /9 6 1 /2 3 /9 6 1 /2 4 /9 6 1 /2 5 /9 6 1 /2 6 /9 6 1 /2 7 /9 6 1 /2 8 /9 6
D a te

Figure 1-18 Strain Readings at Girder 2 During Winter


Temperature Distribution @ Girder 2
100
Bridge Temp
90
Microstrain

80

70

60
Ambient Temp

50
8/24/96 8/25/96 8/26/96 8/27/96 8/28/96 8/29/96 8/30/96 8/31/96
Date
D a ily S tra in C h a n g e s @ G ird e r 2 T o p F la n g e
80
ABUT 2
QSPAN 2
60 P IE R 2
M ID S P A N 2

40
Microstrain

20

-2 0

-4 0

-6 0
D a ily S tr a in C h a n g e s @ G ir d e r 2 B o tto m F la n g e
80

60

40
Microstrain

20

-2 0
ABUT 2
-4 0 QSPAN 2
P IE R 2
M ID S P A N 2
-6 0
8 /2 4 /9 6 8 /2 5 /9 6 8 /2 6 /9 6 8 /2 7 /9 6 8 /2 8 /9 6 8 /2 9 /9 6 8 /3 0 /9 6 8 /3 1 /9 6
D a te

Figure 1-19 Strain Readings at Girder 2 During Summer


Daily Curvature Changes in Winter
4 50

40
3
30
Curvature (10e-6/in)

2 Sign Convention 20
+-

Temp (F)
10
1
0

0 -10

-20
-1 ABUT2 QSPAN2
PIER2 MSPAN2 -30
Amb. Temp
-2 -40
1/21/96 1/22/96 1/23/96 1/24/96 1/25/96 1/26/96 1/27/96 1/28/96
Date

Figure 1-20 Daily Curvature Changes at Girder 2 During Winter

Daily Curvature Changes in Summer


4 100

90
3 80

70
Curvature (10e-6/in)

2
60
Temp (F)
50
1
40

30
0
20

-1 10
ABUT2 QSPAN2
PIER2 MSPAN2
0
Amb. Temp.
-2 -10
8/24/96 8/25/96 8/26/96 8/27/96 8/28/96 8/29/96 8/30/96 8/31/96
Date

Figure 1-21 Daily Curvature Changes at Girder 2 During Summer


70
60 Bridge Temp (F)
Temperature (F)

50 Ambient Temp. (F)


40
30
20
10
0
∆T=76.58 F
MENT

-10

AN
PIER

MIDSP
300
ABUT

250
∆ε =287 µε
200
Microstrain

150

100

50

0
2/3/96 2/5/96 2/7/96 2/9/96 2/11/96 2/13/96 2/15/96
-50
Date

Figure 1-22 Maximum Temperature Change/Week Encountered During Monitoring


[i] Burke, M.P., “Integral Bridges”, Transportation Research Record, No 1275, p.53-61, National
Academy Press Washington, D.C., 1990.

[ii] Soltani, A.A., Kukreti, A.R., “Performance Evaluation of Integral Abutment Bridges”,
Transportation Research Record, No 1371, p.17-25, National Academy Press Washington, D.C.,
1992.

[iii] Burke, M.P., “Integral Bridges: Attributes and Limitations”, Transportation Research
Record, No 1393, p.1-8, National Academy Press Washington, D.C., 1993.

[iv] “High Performance Steel Bridge Concepts”, American Iron and Steel Institute, November
1996.

[v] Burke, M.P., “Semi-Integral Bridges: Movements and Forces”, Transportation Research
Record, No 1460, p.1-7, National Academy Press Washington, D.C., 1994.

[vi] Elgaaly, M, Sandford, T.C., Colby, C., “Testing an Integral Steel Frame Bridge”,
Transportation Research Record, No 1371, p.75-82, National Academy Press Washington, D.C.,
1992.

[vii] Steiger, D.J., “Jointless Bridges Provide Fuel for Controversy” Roads & Bridges,Vol. 31,
No.11, November 1993.

[viii] Sotiropoulos, S. N., Gangarao, H.V.S., “Design Anomalies in Concrete Deck-Steel Stringer
Bridges”, Transportation Research Report, No 1393, p. 31-38, National Academy Press
Washington, D.C., 1993.

Vous aimerez peut-être aussi