You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/274409902

Vibration Versus Deflection Control for Bridges with High-Performance Steel


Girders

Article  in  Transportation Research Record Journal of the Transportation Research Board · December 2011
DOI: 10.3141/2251-03

CITATIONS READS
4 669

4 authors, including:

Hani H Nassif Mayrai Gindy


Rutgers, The State University of New Jersey University of Rhode Island
75 PUBLICATIONS   685 CITATIONS    12 PUBLICATIONS   294 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Fiber-Reinforced Self-Consolidating Concrete for Repair of Bridge Sub-Structures and Fiber-Reinforced Super-Workable Concrete for Infrastructure Construction View
project

Impact of Overweight Trucks on Infarstructrue in New Jersey View project

All content following this page was uploaded by Hani H Nassif on 27 January 2016.

The user has requested enhancement of the downloaded file.


Vibration Versus Deflection Control
for Bridges with High-Performance
Steel Girders
Hani Nassif, Ming Liu, Dan Su, and Mayrai Gindy

The use of high-performance steel (HPS) in highway bridges has proved control, as specified in the AASHTO Load and Resistance Factors
successful in structural performance and in the cost-efficiency of the Design (LRFD) Bridge Design Specifications (1), Subsection 2.5.2.6,
constructed bridges. However, the use of optional deflection criteria as and Table 2.5.2.6.3-1, Traditional Minimum Depths for Constant
specified in a subsection of the AASHTO LRFD Bridge Design Specifications Depth Superstructures, may impede the use of HPS in bridge con-
may impede the use of HPS in highway bridges. Besides the deflection struction. If the deflection requirements become the controlling design
criteria, the current AASHTO specifications provide a depth-to-span issue, the most economical bridge designs, on the basis of ultimate
limitation table for steel superstructure designs. The values in that strength criteria, cannot be achieved. As a result, the rationality of
table are based primarily on the use of Grade 36 steel and were initially using deflection requirements must be carefully studied for current
a carryover from railroad bridge construction. Therefore, both the bridge design practices. One of the important questions that need to
deflection criteria and the depth-to-span limitation need to be evaluated be addressed is whether the vibration can be effectively controlled
for bridges constructed with HPS. This paper presents an investigation by applying the deflection limit specifications to the design. The
of the vibration control (e.g., acceleration and velocity) of HPS bridges use of a deflection design parameter is an option that many states
using a three-dimensional (3-D) dynamic computer model. The 3-D (e.g., New Jersey) impose to control vibration indirectly. Vibration
dynamic model was validated with the use of field test data on various control of a constructed bridge is desired because the service life of
bridges, including a three-span continuous steel girder bridge. A suite a bridge that is too flexible becomes questionable. Furthermore, a
of typical bridges designed with various slab thicknesses and span-to- motorist’s perception that he or she is using a safe bridge is important
depth ratios was selected for this study. In particular, the effects of the to state highway agencies.
steel girder depth and concrete slab thickness on bridge vibration were
In this paper, a correlation between live load–induced vibration
identified. The analysis results indicated that bridge vibrations were
and various HPS bridge parameters is established. An analysis model,
better controlled with the choice of optimal concrete slab thicknesses
based on the dynamic interaction between the bridge, the vehicle, and
(i.e., by adding to the mass and moment of inertia of a composite girder)
the road roughness system, was validated by using experimental data
rather than with specifications of the span-to-depth ratio limits, deflection
from various types of bridges. A suite of HPS bridges were designed
limits, or first natural frequency.
and then analyzed to identify the most sensitive parameters that
would affect bridge vibration and deflection response. Various types
High-performance steel (HPS) was developed in the early 1990s to of truck loading patterns (e.g., single, side by side) and typical road
provide an enhanced structural steel material of Grade 70 yield strength roughness profiles were considered. The resulting bridge responses
with improved welding capabilities and weathering durability. (i.e., acceleration and deflection) were compared to human perception
The use of HPS-485W (70W) steel in highway bridges has proven of vibration. The deflection criteria and depth-to-span (D:L) ratio
to be successful in structural performance and in the cost efficiency specified in the AASHTO LRFD Bridge Design Specifications, as
of the constructed bridges. However, it has been realized that the well as Canadian and Australian code specifications, were also
use of the optional design criteria for deformation (or deflection) evaluated. The effects of the steel girder depth and concrete slab
thickness on bridge vibration were identified. Results showed that
bridge vibration could be effectively controlled by increasing the
H. Nassif, Department of Civil Engineering, College of Engineering, Kyung Hee
thickness of the concrete deck slab (i.e., adding mass and moment
University; Center for Advanced Information Processing, Rutgers Infrastructure
Monitoring and Evaluation Group, Rutgers University, 96 Frelinghuysen Road, of inertia) while maintaining shallow depths for the HPS girders.
Piscataway, NJ 08854. M. Liu, Technical Service Center, Bureau of Reclamation, Moreover, increasing slab thickness improved the durability of the
U.S. Department of the Interior, Denver Federal Center, Building 67 (86-68110), bridge deck.
P.O. Box 250007, Denver, CO 80225-0007. D. Su, Center for Advanced Infor-
mation Processing, Rutgers Infrastructure Monitoring and Evaluation Group,
Rutgers University, 96 Frelinghuysen Road, Piscataway, NJ 08854. M. Gindy,
Department of Civil and Environmental Engineering, University of Rhode Island, BACKGROUND
1 Lippitt Road, 201 Bliss Hall, Kingston, RI 02881. Corresponding author: D. Su,
sulevey@eden.rutgers.edu. Use of deflection limits in the design of highway bridges, particularly
when HPS is desired, is an important consideration in the design
Transportation Research Record: Journal of the Transportation Research Board,
No. 2251, Transportation Research Board of the National Academies, Washington,
process. The AASHTO LRFD Bridge Design Specifications establish
D.C., 2011, pp. 24–33. the use of a deflection design as an optional criterion (1). A number
DOI: 10.3141/2251-03 of state highway agencies specify their own deflection limits as a

24
Nassif, Liu, Su, and Gindy 25

mandatory requirement. New Jersey defaults to AASHTO’s deflection 4. The deflection limit of L/800 was established in the 1930s,
criteria of L/800 for vehicular bridges and L/1,000 for pedestrian primarily for vibration control of steel highway bridges.
bridges (2). It is perceived that these deflection limits are usually 5. No major changes have been made for steel highway bridges
based on a rather arbitrary and sometimes conservative approach since 1936.
in designing bridges (3). This perception raises questions about
the rationality of these deflection limits in that pursuing a conser- It is generally agreed that deformation requirements are intended to
vative design approach does not take advantage of today’s bridge play an important role in bridge vibration controls. Therefore, Wright
construction technologies. In principle, deflection limits were estab- and Walker (7) proposed a vibration-related static deflection limit,
lished to eliminate damage to any structural and nonstructural a computed transient peak acceleration of a bridge, α, which should
components due to excessive deformation as well as to avoid loss not exceed 100 in./s2 (2.54 m/s2), where the static deflection, δs, is
of aesthetic appearance and interruption of its functionality. For linked to α as follows:
highway bridges, vehicle rideability and human responses to bridge
vibration under normal traffic conditions play important roles in α
δs = 0.05 × L × (1)
determining deflection limits. Moreover, if deflection limits should (speed + 0.3 × fs × L ) × fs
be specified as a design parameter, a rational approach is needed
to establish a design philosophy for limit states that all designers where
can justifiably use (4).
A detailed literature search by Gindy (4) indicated that the origin L = span length,
of the deformation requirements in bridge designs might be traced speed = vehicle speed, and
to 1871 with a set of specifications established by the Phoenix Bridge fs = natural frequency of a simple span bridge, which may be
Company. An ASCE report on deflection limits of bridges has been estimated as
widely cited for the evolution of these requirements (5). The following
conclusions may be drawn from the ASCE report: π Eb I b
fs = (2)
2 L2 m
1. The deflection limits were established before the span-to-depth
ratio limits; that is, the span-to-depth ratio limits were used as indirect where EbIb is the flexural rigidity of a girder section, and m is the unit
measures of the deflection limits. mass of a girder section. The value for δs needs to be calculated for
2. Both deflection and span-to-depth ratio limits were empiri- a live load with a girder distribution factor of 0.7. Nowak and
cally derived, mainly for early bridge structures, such as wood Grouni (8) also stated that the deflection and vibration criteria
plank decks, pony trusses, simple rolled beams, and pin-connected should be derived by considering human reactions rather than struc-
through-trusses (6). tural performance. In addition, similar efforts that relate static
3. The specified span-to-depth ratio limits for highway bridges deflection limits to the natural frequency of a bridge are shown in
(by AASHO) followed what was established for railroad bridges Canadian and Australian bridge design specifications (9–12). The
(by the American Railway Engineering Association). red solid lines in Figure 1 show the relationship between the first

220
CSA-w/o sidewalk
200

180 CSA-with sidewalk, little pedestrian use

160 CSA-with sidewalk, significant pedestrian use


Static Deflection (mm)

140
AUSTRALIAN
120

100

80
U
60 U

40
A
20
A
0
0 2 4 6 8 10
First Flexural Frequency (Hz)

FIGURE 1 Static deflection versus first flexural frequency (A ⴝ acceptable, U ⴝ unacceptable, CSA = Canadian
Standards Association) (9–12).
26 Transportation Research Record 2251

flexural frequency (in Hertz) and static deflections limits (in milli- a heavy truck) may not be sufficient to represent anticipated bridge
meters) at the sidewalk or edge of a bridge for various structure vibrations. This is why some bridges that were designed to satisfy a
types from the Ontario Highway Bridge Design Code. This relation- deflection limit may still have objectionable vibrations or unaccept-
ship was developed from extensive field data and analytical models able structural performances while other bridges that fail to meet the
(13). The green solid line in Figure 1 presents a similar relationship existing deflection limits perform satisfactorily.
that is adopted in the Australian Bridge Design Code. The static
deflection limits are usually absent in European bridge design spec-
ifications whereas the New Zealand Bridge Manual in 1994 limits EXPERIMENTAL PROGRAM
the maximum vertical velocity to 2.2 in./s (0.056 m/s) instead of
using the maximum acceleration. The Doremus Avenue Bridge, near the port area of Newark, New
Debates on the necessity of deformation requirements in current Jersey, was selected for the field study. The data collected from this
bridge design specifications focus on two aspects: (a) whether bridge were used to validate a dynamic model developed by Nassif
excessive deflections cause structural damage and (b) whether et al. (24). The dynamic model was used thereafter to analyze various
deflection limits provide effective control of bridge vibrations under bridges designed with HPS. The field data collection was performed
normal truck traffic. The limited survey in the ASCE report (5) as part of an extensive long-term monitoring program (25). The three
revealed no evidence of serious structural damage due to excessive equal spans [45 m (148 ft)] of this bridge were selected for the
vertical deflections. However, unfavorable psychological reactions experimental study because Span 3 provided clear access to the under-
to bridge vibrations caused greater concerns about bridge safety. side of the bridge. Details of the bridge cross section are presented
Burke (14) argued that, if deflection limits were not mandated, the in Figure 2.
effective service life of reinforced concrete deck slabs could become The bridge was instrumented with several types of sensors,
considerably less than their normal replacement interval of 30 years. including a weigh-in-motion system, accelerometers, strain gauges,
Moreover, the report on NCHRP Project 20-7 (15) found that and two types of displacement-measuring systems [a permanent
bridges clearly suffered severe structural damage caused by exces- linear variable differential transformer cable system and a portable
sive deformation but provided little support for the idea that deflec- laser Doppler vibrometer system]. Strain transducers were installed
tion limits should be used as a method of controlling these structural across all 10 girders for all three spans at locations of maximum
damages. In addition, deflection limits were not considered as the positive moments while the linear variable differential transformer
good method of controlling bridge vibrations (16). and laser Doppler vibrometer systems measured the response of
Factors related to human perceptions of vibrations are based on Girder 8 of Span 3 at the same location. The weigh-in-motion system,
research work from many sources (13, 17–23). In addition, compre- installed at the bridge entrance, continuously recorded truck traffic
hensive research on human responses to bridge vibrations was information such as axle weights and spacing, vehicle speed, vehicle
conducted by Wright and Walker (7 ). They concluded that peak class, and lane position. The linear variable differential transformer
accelerations were preferable to peak velocities when human per- system measured the deflection of the girder at the location of the
ceptions of bridge vibrations that typically ranged from 1 to 10 Hz maximum positive moment. The noncontact laser Doppler vibrometer
were being evaluated. Thus, peak acceleration and frequency of provided readings for the velocity and displacement of the girder.
bridge vertical vibrations were considered the most important param- During data collection, when the weight of a truck was greater than
eters in this research. the weight of an HS-20 truck (i.e., 72 kips), this truck was identified
The dynamic characterization of a heavy truck—such as its mass, as a heavy truck, and then the data collection system was triggered
speed, traveling lane, suspension properties, and tire stiffness— to record the bridge response for that loading event. The output files
significantly affects the dynamic response of a bridge. In other words, were all linked by a time stamp and truck identification number.
the predicted dynamic behavior of a bridge under normal moving Periodically, the network was accessed remotely and the files were
loads (i.e., without consideration of the dynamic characterization of downloaded and analyzed.

FIGURE 2 Cross section of instrumented span of the Doremus Avenue Bridge


(SHLDR ⴝ shoulder; SB ⴝ southbound; NB ⴝ northbound; LVDT ⴝ linear variable
differential transformer; LDV ⴝ laser Doppler vibrometer; G ⴝ girder;
Acc ⴝ accelerometer; Geo ⴝ geophone).
Nassif, Liu, Su, and Gindy 27

BRIDGE DYNAMIC MODEL 3


2.5 Field Data
Deformation requirements are intended to play a significant role in 2
3-D Model

Stress (ksi)
controlling bridge vibrations. Therefore, it is important to understand 1.5
the dynamic–vibration behavior of a bridge subjected to normal truck 1
traffic loading conditions. In this study, a three-dimensional (3-D) 0.5
dynamic computer model developed by Nassif et al. (24) was used 0
to predict the responses of girder bridges. The model consisted of -0.5
three submodels—bridge model, vehicle model, and road roughness -1
100 200 300 400 500 600
model—to represent the interaction between the truck, the bridge,
Front Axial Distance from Bridge Entrance (ft)
and the road roughness. The bridge model was based on efficient
grillage analysis. The five-axle tractor–semitrailer vehicle model is FIGURE 3 Validation of 3-D bridge dynamic model by means
a 3-D representation of the most common truck on the highways, of measured stresses at Span 2, Girder 9.
as observed by Nassif (26). The bridge response was computed from
the 3-D computer model by using the Newmark-β method. Nassif
et al. (24) presented the experimental validation of this 3-D model by stresses with the calculated ones from the 3-D model, as shown in
comparing the computed dynamic load factor with the experimental Figure 3. The possible reasons for the variation in the measurements
results on the basis of field tests done by Nassif and Nowak (27). included the road roughness, variation in truck dynamics (e.g., suspen-
As described earlier, the experimental data collected from the sion and axle configuration), and variation of truck-loading location.
Doremus Avenue Bridge were used to validate the model. The 3-D Next, the mass matrix of the bridge model was calibrated by matching
dynamic model was also calibrated by using data from dynamic tests the first natural frequency to the free-vibration testing (i.e., 1.59 Hz,
performed on more than four bridges in the state of Michigan. The as shown in Figure 4) with those calculated from the 3-D model.
Doremus Avenue Bridge grillage model was assembled by using As a result, the calibrated dynamic models were applied to a suite
480 longitudinal and 441 transverse elements. The stiffness matrix of designed HPS bridges with various D:L ratios and slab thicknesses.
of the bridge model was calibrated by matching the measured static A parametric study was performed to identify the effects of the

(a)

(b)

FIGURE 4 Measured acceleration records on Span 3, Girder 8 (interior), middle span (4).
28 Transportation Research Record 2251

14’-8” 4’-7” 23’-4” 4’-6” Truck A

10.76 kips 15.74 15.74 17.05 17.05


kips kips kips kips

13’-4” 4’-5” 23’-6” 4’-7” Truck B

10.34 kips 13.87 13.87 19.26 19.26


kips kips kips kips

FIGURE 5 Configurations of Trucks A and B.

superstructure, truck traffic, and road roughness profiles on bridge from 8.66 to 12 in. (0.220 to 0.305 m), when the steel girder depth
vibration levels, with particular emphasis on the effects of the steel decreases, the maximum static deflections slightly increase. In contrast,
girder depth and concrete slab thickness. Two five-axle trucks with Table 1 shows that the maximum vertical accelerations depend on both
typical truck properties obtained from the field testing of Doremus bridge and truck properties. For example, Truck A generally induces
Avenue Bridge were used in this study. The configurations of these higher vibrations than Truck B for the same bridge. The presence of
two trucks are shown in Figure 5. Both multiple-truck configurations multiple trucks (i.e., side by side) often does not increase bridge
(i.e., two trucks side by side) and one single-truck configuration vibrations, although presence of multiple trucks always generates
were also considered in this study. more static deflection. Figure 7 shows that bridge vibration can be well
controlled under the “perceptible” level by changing the concrete slab
thickness, even if the D:L limit is not satisfied. This result indicates that
ANALYSIS RESULTS a shallower HPS girder can be used if the D:L ratio limit is ignored.
In contrast, if the D:L ratio limit is considered, a larger girder depth
As stated earlier, the deflection limits and the D:L ratio specifications needs to be used, and therefore the cost of the section will increase.
from AASHTO LRFD Bridge Design Specifications needed to be Moreover, Figure 8 shows the acceleration records for two bridges
evaluated. Figure 6 shows a plot of the maximum static deflections designed with the same concrete slab thickness of 8.66 in. (0.220 m)
versus the D:L ratios for different concrete slab thicknesses. The D:L but D:L ratios of 0.016 and 0.032, respectively. In Figure 8, the peak
limitation used was 0.027 for these designed bridges. Figure 6 acceleration decreases by only 16% [from 24.5 in./s2 (0.62 m/s2) to
shows that the deflection limits can still be met even if the minimum 20.49 in./s2 (0.52 m/s2)], even when the D:L ratio is doubled. This result
D:L ratios are not satisfied. For concrete slab thicknesses ranging indicates that the change in the steel girder depth has little effect on

1.2 0.030
t = 8.66"
1.1 t = 10"
D/L < 0.027
t = 12"
Maximum Static Deflection (in.)

Maximum Static Deflection (m)

1.0 0.025
>L / 1000
0.9

0.8 0.020
< L / 1000
0.7

0.6 0.015

0.5
D/L > 0.027
0.4 0.010
0.01 0.02 0.03 0.04
Depth-to-Span Ratio (D/L)

FIGURE 6 Maximum static deflection versus D:L ratio (t ⴝ concrete slab thickness).
Nassif, Liu, Su, and Gindy 29

TABLE 1 Maximum Vertical Acceleration for Different Depth-to-Span Ratios


and Slab Thickness Under Various Truck Loading Conditions

Maximum Acceleration, m/s2 (in./s2)

Truck I II III IV

Concrete Slab Thickness: 8.66 in. (0.220 m)


A 1.39 (54.79) 1.49 (58.64) 1.37 (53.82) 1.42 (55.75)
B 0.91 (35.79) 1.05 (41.23) 0.94 (37.02) 1.46 (57.42)
A and B (side by side) 1.51 (59.60) 1.52 (59.91) 1.40 (55.21) 1.41 (55.53)
Concrete Slab Thickness: 10 in. (0.254 m)
A 1.00 (39.30) 1.01 (39.91) 1.02 (40.29) 1.16 (45.80)
B 0.77 (30.13) 1.05 (41.46) 0.89 (34.99) 0.84 (33.26)
A and B (side by side) 1.14 (45.05) 1.15 (45.08) 1.18 (46.47) 1.18 (46.36)
Concrete Slab Thickness: 12 in. (0.305 m)
A 0.71 (28.09) 0.73 (28.84) 0.80 (31.47) 0.86 (33.74)
B 0.73 (28.86) 0.69 (27.23) 0.76 (29.89) 0.63 (24.80)
A and B (side by side) 0.71 (28.14) 0.73 (28.55) 0.73 (28.89) 0.77 (30.24)

NOTE: I: steel girder depth, m (in.) = 1.44 (56.69); D:L = 0.032; satisfies minimum D:L of 0.027.
II: steel girder depth, m (in.) = 1.20 (47.24); D:L = 0.027; satisfies minimum D:L of 0.027.
III: steel girder depth, m (in.) = 0.96 (37.79); D:L = 0.021; does not satisfy minimum D:L of 0.027.
IV: steel girder depth, m (in.) = 0.72 (28.34); D:L = 0.016; does not satisfy minimum D:L of 0.027.

bridge vibrations for the same concrete slab thickness. In contrast, the static deflections calculated in accordance with the New Jersey
Figure 9 shows the acceleration records for two bridges having the bridge design specifications and those specified in the Australian
same D:L ratio but slab thicknesses of 8.66 in. (0.220 m) and 12 in. and Canadian bridge design codes. Although all 12 of these bridges
(0.305 m). When the slab thickness increases by 38.5% [from 8.66 in. met the maximum static-deflection limits for the Australian bridge
(0.220 m) to 12 in. (0.305 m)], the peak acceleration decreases by 54% design codes, all but two bridges [those having a concrete slab thick-
[from 35.72 in./s2 (0.91 m/s2) to 16.56 in./s2 (0.42 m/s2)]. This result ness of 8.66 in. (0.220 m) and D:L ratios equal to 0.027 and 0.032,
suggested that the increase in slab thickness had a significant effect respectively] meet the maximum static deflection limits in the
on the vibration control of HPS bridges. Canadian bridge design code. Moreover, although two bridges
As mentioned earlier, vehicle rideability (or road roughness) and [those having concrete slab thickness of 8.66 in. (0.220 m) and D:L
human responses to bridge vibration play important roles in deter- ratios of 0.016 and 0.022, respectively] do not meet the New Jersey
mining deflection limits. That is why the Australian and Canadian bridge design specifications (i.e., deflection limit of L/1,000), both
bridge design codes relate the static deflection limits to the first natural are within the maximum static deflection limits of the Canadian and
frequency of the bridge, as shown earlier in Figure 1. Figure 10 shows Australian bridge design specifications. Figures 11a through 11c

70
t = 8.66" 1.7
Maximum Vertical Acceleration (in./sec.2)

Maximum Vertical Acceleration (m/sec.2)

D/L < 0.027 t = 10"


60 t = 12"
1.5
Unpleasant

1.3
50

1.1
Perceptible
40
0.9

30
0.7
D/L > 0.027

20 0.5
0.01 0.02 0.03 0.04
Depth-to-Span Ratio (D/L)

FIGURE 7 Maximum vertical acceleration versus D:L ratio.


30 Transportation Research Record 2251

30 0.8
t = 8.66", D/L=0.016
0.6
20
t = 8.66", D/L=0.032
0.4

Acceleration (in./sec.2)

Acceleration (m/sec.2)
10
0.2

0 0

-0.2
-10
-0.4
-20
-0.6

-30 -0.8
0 1 2 3 4 5 6 7 8 9 10 11 12
Time (sec)

FIGURE 8 Acceleration records with identical slab thickness.

30
t = 8.66" 0.6
20 t = 12" 0.4
Acceleration (in./sec.2)

Acceleration (m/sec.2)
10 0.2

0 0.0

-0.2
-10
-0.4
-20
-0.6

-30 -0.8
0 1 2 3 4 5 6 7 8 9 10 11 12
Time (sec)

FIGURE 9 Acceleration records with different slab thickness.

1.2 0.030
t = 8.66"
t = 10" Australian
1.1
bridge design
Maximum Static Deflection (in.)

Maximum Static Deflection (m)

t = 12" code
1.0 0.025
> L/1000
New Jersey bridge
0.9
design specifications
0.8 0.020
< L/1000
0.7
Canadian
0.6 bridge design 0.015
code
0.5

0.4 0.010
1.2 1.3 1.4 1.5 1.6 1.7 1.8
First Natural Frequency (Hz)

FIGURE 10 Maximum static deflection versus first natural frequency.


Nassif, Liu, Su, and Gindy 31

70
Truck A 1.7

Maximum Vertical Acceleration (in./sec.2)

Maximum Vertical Acceleration (m/sec.2)


Truck B
60 1.5
Trucks A & B Unpleasant
(Side by Side) 1.3
50

1.1
40
0.9
Perceptible*
30
0.7

20 0.5
1.2 1.3 1.4 1.5 1.6 1.7 1.8
First Natural Frequency (Hz)
(a)

70
Truck A 1.7

Maximum Vertical Acceleration (m/sec.2)


Maximum Vertical Acceleration (in./sec.2)

Truck B
60 1.5
Trucks A & B (Side Unpleasant
by Side) 1.3
50

1.1
40
0.9
Perceptible*
30
0.7

20 0.5
1.2 1.3 1.4 1.5 1.6 1.7 1.8
First Natural Frequency (Hz)
(b)

70
Truck A 1.7
Maximum Vertical Acceleration (in./sec.2)

Maximum Vertical Acceleration (m/sec.2)

Truck B
60 1.5
Trucks A & B Unpleasant
(Side by Side) 1.3
50

1.1
40
0.9
Perceptible*
30
0.7

20 0.5
1.2 1.3 1.4 1.5 1.6 1.7 1.8
First Natural Frequency (Hz)
(c)

FIGURE 11 Maximum vertical acceleration versus first natural frequency for concrete
slab thickness of (a) 8.66 in., (b) 10 in., and (c) 12 in. (perceptible vibration calculated
as 0.5 ⴛ frequency ).
32 Transportation Research Record 2251

Moment of Inertia (m4)


0.17 0.27 0.37 0.47 0.57 0.67
70
t = 8.66" 1.7
Unpleasant

Maximum Vertical Acceleration (in./sec.2)

Maximum Vertical Acceleration (m/sec.2)


t = 10"
60 1.5
t = 12"

1.3
50

1.1
Perceptible
40

0.9

30
0.7

20 0.5
20 40 60 80
Moment of Inertia (ft4)

FIGURE 12 Variation of maximum vertical acceleration versus composite girder moment


of inertia.

show plots of the maximum vertical accelerations versus the first parameters to control the vibration of steel girder bridges. From the
natural frequency for different concrete slab thicknesses under analysis results, the following conclusions can be made:
various loading conditions. The bridges with a concrete slab thick-
ness of 8.66 in. (0.220 m) experience the most unpleasant vibrations • Neither D:L limitations nor static deflection criteria could pro-
(Figure 11a). Moreover, Figures 11a through 11c provide strong vide effective vibration control of a steel girder bridge under normal
evidence that the bridges with the same first natural frequency may truck traffic conditions.
experience large variations in their maximum vertical accelerations • If the D:L ratio limit is not considered, the more economical
under different truck loads. In other words, the first natural frequency, design can be achieved by choosing shallower HPS girder sections
as used by the Canadian as well as Australian codes, does not effec- while controlling the bridge vibration well under the perceptible level
tively control bridge vibrations. Figure 12 shows the variation in the by increasing concrete slab thickness.
maximum acceleration versus the girder moment of inertia. A strong • Bridges with the same first natural frequency may experience
correlation exists between maximum acceleration and girder moment large variations in their maximum vertical accelerations under
of inertia. Figure 12 shows the effect of slab thickness in reducing different truck loads. Therefore, the first natural frequency of the
maximum vertical acceleration, indicating that control of vibration bridge, as used by the Canadian and Australian codes, may not
can also be achieved by increasing mass and stiffness of the com- effectively control bridge vibrations.
posite girder without the need to increase steel girder depth. The • The vertical acceleration of HPS girder bridges can be effectively
moment of inertia of the composite girder increased by only 10% controlled by increasing the mass and stiffness of the composite
(from 51.6 to 57.3 ft4) when the D:L increased by 52% (from 0.021 girder (i.e., by increasing concrete deck thickness) without the need
to 0.032) compared with an increase of 22% (from 51.6 to 66.5 ft4) to increase HPS girder depth.
when the slab thickness increased by 20% (from 10 to 12 in.) with
a D:L equal to 0.021. These data confirm that the increase in slab
thickness can enhance the stiffness of the composite girder more ACKNOWLEDGMENTS
effectively and lead to a further reduction in the vibration of the
structure. The authors gratefully acknowledge the New Jersey Department of
Transportation for sponsoring the research project related to instru-
mentation and monitoring of the Doremus Avenue Bridge. They
CONCLUSIONS also acknowledge the financial support and the technical assistance
of Harry Capers, Jr. (retired), Nick Vittilo (retired), Jose Lopez
This paper presented results of a comprehensive evaluation of the (retired), W. Lad Szalaj (retired), and Dick Dunne of the New Jersey
deflection criteria from various codes by using a 3-D dynamic model. Department of Transportation and the help of Nakin Suksawang,
A parametric study was conducted to identify the most sensitive Joe Davis, and Talat Abu-Amra.
Nassif, Liu, Su, and Gindy 33

REFERENCES 2002. http://www.trb.org/Main/Blurbs/Improved_Live_Load_Deflection_


Criteria_for_Steel_B_161461.aspx. Accessed Sept. 1, 2008.
1. AASHTO LRFD Bridge Design Specifications. AASHTO, Washington, 16. Azizinamini, A., K. Barth, R. Dexter, and C. Rubeiz. High Performance
D.C., 2008. Steel: Research Front, Historical Account of Research Activities.
2. New Jersey Bridge Design Specifications. New Jersey Department of Journal of Bridge Engineering, Vol. 9, No. 3, 2004, pp. 212–217.
Transportation, Trenton, 2006. 17. Reiher, H. and F. J. Meister. The Effect of Vibration on People
3. Demitz, J. R., D. R. Mertz, and J. W. Gillespie. Deflection Requirements [Forschung auf dem Gebeite des Ingenieurwesens (in German)]. Report
for Bridges Constructed with Advanced Composite Materials. Journal F-TS-616-RE, Vol. 2, No. II, P381. Headquarters Air Material Command,
of Bridge Engineering, Vol. 8, No. 2, March 2003, pp. 73–83. Wright Army Air Field, Ohio, 1946.
4. Gindy, M. Development of a Reliability-Based Deflection Limit State 18. Goldman, D. E. A Review of Subjective Responses to Vibratory Motion
for Steel Girder Bridges. PhD dissertation. Rutgers, State University of of the Human Body in the Frequency Range 1 to 70 Cycles per Second.
New Jersey, Piscataway, 2004. Naval Medical Research Institute, National Naval Medical Center,
5. Committee on Deflection Limitations of Bridges. Deflection Limitation Bethesda, Md., 1948.
of a Bridge (Report ST3). Journal of the Structural Division of ASCE, 19. Janeway, R. N. Vehicle Vibration Limits for Passenger Comfort: From
Vol. 84, May 1958, pp. 1633–1720. Ride and Vibration Data. SP-6. Society of Automotive Engineers, Detroit,
6. Wu, H. Influence of Live-Load Deflections on Superstructure Performance Mich., 1950.
of Slab on Steel Stringer Bridges. PhD dissertation. West Virginia 20. Oehler, L. T. Vibration Susceptibilities of Various Highway Bridge Types.
University, Parkersburg, 2003. Project 55 F-40, No. 272. Michigan State Highway Department, Lansing,
7. Wright, R. N., and W. H. Walker. Criteria for the Deflection of Steel 1957.
Bridges. Bulletin of the American Iron and Steel Institute, No. 19, 21. Oehler, L. T. Highway Research Circular 107: Bridge Vibration:
Nov. 1971. Summary of Questionnaire to State Highway Departments. HRB, National
8. Nowak, A. S., and H. N. Grouni. Serviceability Considerations for Research Council, Washington, D.C., Feb. 1970.
Guideways and Bridges. Canadian Journal of Civil Engineering, Vol. 15, 22. Oriard, L. L. Blasting Operations in the Urban Environment. Bulletin of
No. 4, 1988, pp. 534–537. the Association of Engineering Geologists, Vol. 9, No. 1, 1972, pp. 27–46.
9. Ontario Highway Bridge Design Code: Commentary, 3rd ed. Ministry 23. Ontario Highway Bridge Design Code: Commentary and March 1995
of Transportation, Quality, and Standards Division, Toronto, Ontario, Update. Quality and Standards Division, Ontario Ministry of Trans-
Canada, 1991. portation, Toronto, Ontario, Canada, 1995.
10. CAN/CSA: S6-00 and Commentary. Canadian Highway Bridge Design 24. Nassif, H. H., M. Liu, and O. Ertekin. Model Validation for Bridge–Road–
Code. Canadian Standards Association, Toronto, Ontario, Canada, Vehicle Dynamic Interaction System. Journal of Bridge Engineering,
May 2000. Vol. 8, No. 2, March 2003, pp. 112–120.
11. ’92 Austroads Bridge Design Code, Section Two: Code Design Loads 25. Nassif, H., N. Suksawang, J. Davis, M. Gindy, and T. Abu-Amra.
and Commentary. Austroads, Haymarket, New South Wales, Australia, Instrumentation, Field Testing, and Monitoring of the Doremus Avenue
1992. Bridge: Superstructure. FHWA-NJ-2005-13. New Jersey Department
12. ’96 Australian Bridge Design Code, Section Six: Steel and Composite of Transportation, Trenton, 2010.
Construction. Austroads, Haymarket, New South Wales, Australia, 1996. 26. Nassif, H. H. Live Load Spectra for Girder Bridges. PhD dissertation.
13. Wright, D. T., and R. Green. Highway Bridge Vibration, Part II: Report University of Michigan, Ann Arbor, 1993.
No. 5 on Ontario Test Programme. Ontario Department of Highways 27. Nassif, H. H., and A. S. Nowak. Dynamic Load Spectra for Girder
and Queen’s University. Kingston, Ontario, Canada, May 1964. Bridges. In Transportation Research Record 1476, TRB, National
14. Burke, M. P. Superstructure Flexibility and Disintegration of Reinforced Research Council, Washington, D.C., 1995, pp. 69–83.
Concrete Deck Slabs: An LRFD Perspective. In Transportation Research
Record: Journal of the Transportation Research Board, No. 1770, TRB, The findings reported in this paper are those of the authors and do not necessarily
National Research Council, Washington, D.C., 2001, pp. 76–83. reflect the views of the New Jersey DOT. The analytical work by M. Liu was done before
15. Roeder, C. W., K. Barth, and A. Bergman. NCHRP Web Document 46: his employment at the Bureau of Reclamation, U.S. Department of the Interior.
Improved Live Load Deflection Criteria for Steel Bridges. Transportation
Research Board of the National Academies, Washington, D.C., May The General Structures Committee peer-reviewed this paper.

View publication stats

You might also like