Vous êtes sur la page 1sur 21

Nassif, Liu, Su, and Gindy 1

1 VIBRATION VERSUS DEFLECTION CONTROL FOR HIGH-PERFORMANCE


2 STEEL (HPS) GIRDER BRIDGES
3
4 Hani Nassif, Ph.D., P.E., Associate Director
5 Center for Advanced Information Processing (CAIP)
6 Rutgers Infrastructure Monitoring and Evaluation (RIME) Group
7 Rutgers, The State University of New Jersey
8 96 Frelinghuysen Road, Piscataway, NJ 08854
9 Phone: (732) 445-4414, Fax: (732) 445-4775
10 nassif@rci.rutgers.edu
11
12 Ming Liu, Ph.D., P.E.
13 Civil Engineer (Structural)
14 Technical Service Center - Bureau of Reclamation
15 U.S. Department of the Interior
16 Denver Federal Center
17 Bldg. 67 (86-68110)
18 P.O. Box 25007, Denver, CO 80225-007
19 mliu@usbr.gov
20
21 Dan Su*, Ph.D. Candidate, Research Assistant
22 Center for Advanced Information Processing (CAIP)
23 Rutgers Infrastructure Monitoring and Evaluation (RIME) Group
24 Rutgers, The State University of New Jersey
25 96 Frelinghuysen Road, Piscataway, NJ 08854
26 sulevey@eden.rutgers.edu
27
28 Mayrai Gindy, Ph.D., Associate Professor
29 Dept. of Civil & Environmental Engineering
30 University of Rhode Island
31 1 Lippitt Road, 201 Bliss Hall
32 Kingston, RI 02881
33 gindy@egr.uri.edu
34
35 * Corresponding Author
36 Word count: 4093
37 Abstract: 249< 250
38 Figures & Tables: 13x250 = 3250
39 Total: 7343
40 Submission Date: 11/15/2010
41

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 2

1 ABSTRACT
2 Utilization of High Performance Steel (HPS) in highway bridges has proven to be successful in
3 terms of structural performance and in the cost efficiency of the constructed bridges. However,
4 the use of optional deflection criteria as specified in the American Association of State Highway
5 Transportation Officials (AASHTO) Load and Resistance Factor Design (LRFD) Bridge Design
6 Specifications, Article 2.5.2.6, may impede the use of HPS in highway bridges. Besides the
7 deflection criteria, the current AASHTO LRFD Bridge Design Specifications also provide a
8 depth-to-span limitation table 2.5.2.6.3-1 for steel superstructure designs. The values in table
9 2.5.2.6.3-1 are primarily based on the use of Grade 36 steel and were initially a carryover from
10 railroad bridge construction. Therefore, both the deflection criteria and depth-to-span limitation
11 need to be evaluated for bridges constructed with HPS. This paper presents an investigation of
12 the vibration control (e.g. acceleration and velocity) of HPS bridges using a 3-D dynamic
13 computer model. The 3-D dynamic model was validated using field test data on various bridges
14 including a 3-span continuous steel girder bridge. A suite of typical bridges designed with
15 various slab thicknesses and span-to-depth ratios were selected for this study. In particular, the
16 effects of the steel girder depth and concrete slab thickness on bridge vibration were identified.
17 The analysis results indicated that bridge vibrations is better controlled by choosing optimal
18 concrete slab thicknesses (i.e. adding to the mass and moment of inertia of composite girder)
19 rather than specifying the span-to-depth ratio limits, deflection limits, or first natural frequency.
20
21 Key Words:
22 Vibration
23 Bridges
24 Deflection
25 Slab thickness
26 Span-to-depth ratio
27

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 3

1 INTRODUCTION
2 High Performance Steel (HPS) was developed in the early 1990s to provide an enhanced
3 Grade 70 yield strength structural steel material, improved welding capabilities and weathering
4 durability. The use of HPS-485W (70W) steel in highway bridges has proven to be successful in
5 terms of structural performance and in the cost efficiency of the constructed bridges. However,
6 it has been realized that the use of the optional design criteria for deformation (or deflection)
7 control, as specified in American Association State Highway Transportation Officials
8 (AASHTO) Load and Resistance Factors Design (LRFD) Bridge Design Specifications [1],
9 Subsection 2.5.2.6, and Table 2.5.2.6.3-1 Traditional Minimum Depths for Constant Depth
10 Superstructures may impede the use of HPS in bridge construction. If the deflection
11 requirements become the controlling design issue, the most economical bridge designs, based on
12 ultimate strength criteria, cannot be achieved. As a result, the rationality of using deflection
13 requirements must be carefully studied for current bridge design practices. One of the important
14 questions that needs to be addressed is whether the vibration can be effectively controlled by
15 applying the deflection limit specifications to the design? The use of a deflection design
16 parameter is an option that many states (e.g. New Jersey) impose in order to indirectly control
17 vibration. Vibration control of a constructed bridge is desired because the service life of a bridge
18 that is too flexible becomes questionable. Also, a motorists perception of using a safe bridge is
19 important to the State Highway Agencies.
20 In this paper, a correlation between live load-induced vibration and various HPS bridge
21 parameters is established. An analysis model, based on the dynamic interaction between the
22 bridge, vehicle, and road roughness system, is validated using experimental data from various
23 types of bridges. A suite of HPS bridges are designed and then analyzed to identify the most
24 sensitive parameters that would affect bridge vibration and deflection response. Various types of
25 truck loading patterns (e.g. single, side by side) are considered. Typical road roughness profiles
26 are considered. The resulting bridge responses (i.e. acceleration and deflection) are compared
27 with human perception of vibration. The deflection criteria and depth-to-span ratio specified in
28 AASHTO LRFD Bridge Design Specifications, as well as Canadian and Australian Code
29 Specifications, are also evaluated. The effects of the steel girder depth and concrete slab
30 thickness on bridge vibration are identified. Results show that bridge vibration can be
31 effectively controlled by increasing the thickness of the concrete deck slab (i.e., adding mass and
32 moment of inertia) while maintaining shallow depths for the HPS girders. Moreover, increasing
33 the slab thickness would improve the durability of the bridge deck.
34
35 BACKGROUND
36 Use of deflection limits in the design of highway bridges, particularly when HPS is desired, is an
37 important consideration in the design process. The AASHTO LRFD Bridge Specifications
38 (2008) [1] establishes the use of a deflection design as an optional criteria. A number of State
39 Highway Agencies specify their own deflection limits as a mandatory requirement. New Jersey
40 [2] defaults to AASHTOs deflection criteria of L/800 for vehicular bridges and L/1000 for
41 pedestrian bridges. It is perceived that these deflection limits are usually based on a rather
42 arbitrary and sometimes conservative approach in designing bridges [3]. This raises questions
43 with regard to the rationality of these deflections limits in that pursuing a conservative design
44 approach does not take advantage of todays bridge construction technologies. In principle,

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 4

1 deflection limits were established to eliminate damage to any structural and non-structural
2 components due to excessive deformations as well as to avoid loss of aesthetic appearance and
3 interruption of its functionality. For highway bridges, vehicle rideability and human responses to
4 bridge vibration under normal traffic conditions play an important role in determining the
5 deflection limits. Moreover, if deflection limits should be specified as a design parameter, there
6 is a need for a rational approach to establish a limit state design philosophy that all designers can
7 justifiably use [5].
8 Gindy (2004) [5] performed a detailed literature search that indicated that the origin of the
9 deformation requirements in bridge designs might be traced back to 1871 with a set of
10 specifications established by the Phoenix Bridge Company (PBC). The American Society of
11 Civil Engineering (ASCE) report on deflection limits of bridges (ASCE, 1958) [15] has been
12 widely cited for the evolution of these requirements. The following conclusions may be drawn
13 from the ASCE report:
14 (1) The deflection limits were established before the span-to-depth ratio limits; that is, the
15 span-to-depth ratio limits were used as indirect measures of the deflection limits;
16 (2) Both deflection and span-to-depth ratio limits were empirically derived, mainly for the
17 early bridge structures, such as wood plank decks, pony trusses, simple rolled beams,
18 and pin-connected through-trusses (Wu, 2003) [16];
19 (3) The specified span-to-depth ratio limits for highway bridges (by AASHO) followed
20 what was established for railroad bridges (by American Railway Engineering
21 Association (AREA));
22 (4) The deflection limit of L/800 was established in the 1930s, primarily for the vibration
23 control of steel highway bridges; and
24 (5) No major changes have been made for steel highway bridges since 1936.
25 It is generally agreed that deformation requirements are intended to play an important role in
26 bridge vibration controls. Therefore, Wright and Walker (1971) [12] proposed a vibration-related
27 static deflection limit. The limit is a computed transient peak acceleration of a bridge, , which
28 should not exceed 100 in./sec^2 (2.54 m/sec^2), where the static deflection, s , is linked to as
29 follows:

30 s = 0.05 L (1)
(speed + 0.3 f s L) f s

31 where L = span length; speed = vehicle speed; and fs = natural frequency of a simple span
32 bridge, which may be estimated as

Eb Ib
33 fs = (2)
2L2 m
34 Where Eb I b = flexural rigidity of a girder section; m = unit mass of a girder section. It
35 should be noted that s needs to be calculated for a live load with a girder distribution factor of
36 0.7. Nowak and Grouni (1988) [17] also stated that the deflection and vibration criteria should
37 be derived by considering human reactions rather than structural performance. In addition,
38 similar efforts that relate static deflection limits to the natural frequency of a bridge are shown in

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 5

1 Canadian and Australian bridge design specifications (Ministry, 1991; CSA International, 2000;
2 AUSTROADS, 1992, AUSTRALIAN, 1996) [18], [19], [20], [21]. The red lines in Figure 1
3 shows the relationship between the first flexural frequency (in Hz) and static deflections limits
4 (in mm) at the sidewalk or edge of a bridge for various structure types from the Ontario Highway
5 Bridge Design Code (OHBDC). This was developed from extensive field data and analytical
6 models (Wright and Green, 1964) [11]. The green solid line in Figure 1 presents a similar
7 relationship that is adopted in the Australian Bridge Design Code. It is also noted that the static
8 deflection limits are usually absent in European bridge design specifications while the New
9 Zealand Bridge Manual in 1994 limits the maximum vertical velocity to 2.2 in./sec. (0.056
10 m/sec.) instead of using the maximum acceleration.
11 Debates on the necessity of deformation requirements in current bridge design specifications
12 focus on two aspects: (1) whether excessive deflections cause structural damage; and (2) whether
13 deflection limits provide effective control of bridge vibrations under normal truck traffic. Based
14 on the limited survey in the ASCE report (1958) [15], no evidence of serious structural damage
15 due to excessive vertical deflections was revealed. However, unfavorable psychological
16 reactions to bridge vibrations caused more concerns on bridge safety. Burke (2001) [23] argued
17 that, if deflection limits were not mandated, the effective service life of reinforced concrete deck
18 slabs could become considerably less than their normal replacement interval of 30 years.
19 Moreover, the report on National Cooperative Highway Research Program (NCHRP) project 20-
20 7 (Roeder et al., 2002) [24] found that bridges clearly suffered severe structural damage due to
21 excessive deformation but provided little support for the idea that deflection limits should be
22 used as a method of controlling these structural damages. In addition, deflection limits were not
23 considered as the good method of controlling bridge vibrations (Azizinamini et al., 2004) [25].
24 Factors regarding human perceptions to vibrations are based on research work by Reiher and
25 Meister (1931) [6], Goldman (1948) [7], Janeway (1950) [8], (Oehler, 1957, 1970) [9], [10],
26 Wright and Green (1964) [11], Oriard (1972) [22], and Ontario Ministry of Transportation (1995)
27 [13]. A comprehensive research on human responses to bridge vibrations was conducted by
28 Wright and Walker (1971) [12]. They concluded that peak accelerations were preferable to peak
29 velocities when evaluating human perceptions to bridge vibrations that typically ranged from 1
30 to 10 Hz. Thus, peak acceleration and frequency of bridge vertical vibrations will be considered
31 to be the most important parameters in this research.
32 It should be emphasized that the dynamic characterization of a heavy truck, such as its mass,
33 speed, traveling lane, suspension properties, and tire stiffness, significantly affects the dynamic
34 response of a bridge. In other words, the predicted dynamic behavior of a bridge under normal
35 moving loads (i.e. without consideration of the dynamic characterization of a heavy truck) may
36 not be sufficient to represent anticipated bridge vibrations. This is why some bridges that were
37 designed to satisfy a deflection limit may still have objectionable vibrations or unacceptable
38 structural performances while other bridges that fail to meet the existing deflection limits
39 perform satisfactorily.
40
41 EXPERIMENTAL PROGRAM
42 The Doremus Avenue Bridge, located near the port area of Newark, New Jersey, was selected for
43 the field study. The data collected from this bridge was used to validate a dynamic model
44 developed by Nassif et al. (2003) [26]. The dynamic model was used thereafter to analyze

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 6

1 various bridges designed with HPS. The field data collection was performed as part of an
2 extensive long-term monitoring program [Nassif et. al (2010)] [14]. The three equal spans (45 m,
3 148 ft) of this bridge were selected for the experimental study because Span 3 provides clear
4 access to the underside of the bridge. Details of the bridge cross section are presented in Figure 2.
5 The bridge was instrumented with several types of sensors including a Weigh-In-Motion
6 (WIM) system, accelerometers, strain gauges, and two types of displacement-measuring systems,
7 a permanent Linear Variable Differential Transformer (LVDT)-cable system and a portable laser
8 Doppler Vibrometer (LDV) system. Strain transducers are installed across all 10 girders for all 3
9 spans at maximum positive moment locations while the LVDT and LDV systems measure the
10 response of Girder 8 of Span 3 at the same location. The WIM system, installed at bridge
11 entrance, continuously records truck traffic information such as axle weights and spacing,
12 vehicle speed, vehicle class and lane position. The LVDT system measured the deflection of the
13 girder at the maximum positive moment location. The non-contact LDV provides readings for
14 the velocity and displacement of the girder.
15 During the data collection, when the weight of a truck is larger than the weight of an HS-20
16 truck (i.e., 72 kips), this truck is identified as a heavy truck, and then the data collection system
17 is triggered to record the bridge response for that loading event. The output files are all linked
18 by a timestamp and truck ID number. Periodically, the network is accessed remotely and the
19 files are downloaded and analyzed.
20
21 BRIDGE DYNAMIC MODEL
22 Deformation requirements are intended to play a significant role in controlling bridge vibrations.
23 Therefore, it is important to understand the dynamic/vibration behavior of a bridge subjected to
24 normal truck traffic loading conditions. In this study, a three-dimensional dynamic computer
25 model (3-D Model) developed by Nassif et al. (2003) [26] was used to predict the responses of
26 girder bridges. The model consists of three sub-models: bridge model, vehicle model, and road
27 roughness model to represent the interaction between the truck, bridge, and road roughness. The
28 bridge model is based on the efficient grillage analysis approach. The five axle semi tractor-
29 trailer vehicle model is a three dimensional representation of the most common truck on the
30 highways as observed by Nassif (1993) [4]. The bridge response is computed from the 3-D
31 computer model using the Newmark- method. Nassif et al. (2003) [26] presented the
32 experimental validation of this 3-D model by comparing computed Dynamic Load Factor (DLF)
33 with the experimental results based on field tests done by Nassif and Nowak (1995) [27].
34 As described previously, the experimental data collected from the Doremus Avenue Bridge
35 was used to validate the model. It is also noted that the 3-D dynamic model has also been
36 calibrated using data from dynamic tests performed on more than four bridges in the State of
37 Michigan. The Doremus Avenue Bridge grillage model was assembled using 480 longitudinal
38 and 441 transverse elements. The stiffness matrix of the bridge model is calibrated by matching
39 the measured static stresses with the calculated ones from 3-D Model, as shown in Figure 3. The
40 possible reasons causing the variation of the measurements includes the road roughness,
41 variation in truck dynamics (e.g., suspension and axle configuration), and variation of truck
42 loading location. Next, the mass matrix of the bridge model is calibrated by matching the first
43 natural frequency to the free vibration testing (i.e. 1.59 Hz as shown in Figure 4) with those
44 calculated from 3-D Model.

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 7

1 As a result, the calibrated dynamic models were applied to a suite of designed HPS bridges
2 with various depth-to-span ratio and slab thickness. A parametric study was performed to
3 identify the effects of the superstructure, truck traffic, and road roughness profiles on bridge
4 vibration levels, with particular emphasis on the effects of the steel girder depth and concrete
5 slab thickness. Two 5-axle trucks obtained from the field testing of Doremus Avenue Bridge
6 were used in this study with typical truck properties. The configurations of these two trucks are
7 shown in Figure 5. Both multiple truck configurations (i.e. two trucks side-by-side) and one
8 single truck configuration were also considered in this study.
9
10 ANALYSIS RESULTS
11 As stated earlier, the deflection limits and span-to-depth ratio specifications from AASHTO
12 LRFD Bridge Design Specification need to be evaluated. Figure 6 shows a plot of the maximum
13 static deflections versus the depth-to-span ratios for different concrete slab thicknesses. It is
14 noted that the depth-to-span limitation was used as D/L = 0.027 for these designed bridges.
15 From Figure 6, it can be observed that the deflection limits can still be met, even if the minimum
16 depth-to-span ratios are not satisfied. It is also observed that, for concrete slab thickness ranging
17 from 8.66 to 12 (0.220 m to 0.305 m), when there is a decrease in the steel girder depth there is
18 a slight increase in the maximum static deflections. On the other hand, Table 1 shows that the
19 maximum vertical accelerations depend on both bridge and truck properties. For example, Truck
20 A generally induces higher vibrations than Truck B for the same bridge. The multiple-presence
21 of trucks (i.e. side by side) often does not increase the bridge vibrations, although multiple
22 presences of trucks always generate more static deflection. Figure 7 shows that the bridge
23 vibration can be well controlled under the perceptible level by changing the concrete slab
24 thickness, even if the D/L limit is not satisfied. This indicates that a shallower HPS girder can be
25 used if the depth-to-span ratio limit is ignored. On the other hand, if the depth-to-span ratio limit
26 is considered, a larger girder depth needs to be used, which means that the cost of the section will
27 increase. Moreover, Figure 8 shows the acceleration records in the time domain for the concrete
28 slab thickness of 8.66 inches (0.220 m). It can be observed the peak acceleration decreases from
29 24.5 in./sec.2 (0.62 m/sec.2) to 20.49 in./sec.2 (0.52 m/sec.2), which is around 16 % of reduction,
30 even if the depth-to-span ratio changes from 0.016 to 0.032 (i.e. doubling the depth-to-span ratio).
31 This indicates that the change in the steel girder depth has little effect on the bridge vibrations for
32 the same concrete slab thickness. In contrast, Figure 9 shows the acceleration records in the time
33 domain. It can be observed that when the slab thickness increased from 8.66 in. (0.220 m) to 12
34 in. (0.305 m), the peak acceleration deceased by 54%, from 35.72 in./sec.2 (0.91 m/sec.2) to
35 16.56 in./sec.2 (0.42 m/sec.2). This suggests that the increase in slab thickness has a significant
36 effect on the vibration control of HPS bridges.
37 As mentioned previously, vehicle rideabilities and human responses to bridge vibration play
38 an important role in determining the deflection limits. Therefore, the Australian and Canadian
39 bridge design codes relate the static deflection limits to the first natural frequency of bridge as
40 was shown earlier in Figure 1. Figure 10 shows the calculated static deflections per the New
41 Jersey Bridge Design Specifications and those specified in the Australian and Canadian bridge
42 design codes. It is interesting to note that, while the deflection limits are met according to the
43 Australian bridge design codes for all of these 12 bridges, the two bridges that have a concrete
44 slab thickness of 8.66 (0.220 m) with depth-to-span ratio equal to 0.027 and 0.032, respectively,
45 do not meet the deflection limits in the Canadian bridge design code. Moreover, the two bridges

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 8

1 that do not meet the NJ bridge design specifications (i.e. L/1000) do have the concrete slab
2 thickness of 8.66 (0.220 m) with reduced depth-to-span ratio of 0.016 and 0.022, respectively.
3 Figure 11 (a), (b), and (c) show plots of the maximum vertical accelerations versus the first
4 natural frequency for different concrete slab thicknesses under various loading conditions. The
5 bridges with the concrete slab thickness of 8.66 (0.220 m) experience the most unpleasant
6 vibrations (see Figure 11 (a)). Moreover, Figure 11 provides strong evidence that the bridges
7 with the same first natural frequency may experience large variation in their maximum vertical
8 accelerations under different truck loads. In other words, the first natural frequency, as used by
9 the Canadian as well as Australian Codes, does not effectively control bridge vibrations. Figure
10 12 shows the variation of the maximum acceleration versus the girder moment of inertia. It is
11 observed that there is a strong correlation between maximum acceleration and girder moment of
12 inertia. Figure 12 illustrates the effect of the slab thickness in reducing the maximum vertical
13 acceleration, indicating that control of vibration can be achieved by increasing mass and stiffness
14 of the composite girder without the need to increase steel girder depth. The authors also noted
15 that the moment of inertia of composite girder increased by 10% (from 51.6 ft4 to 57.3 ft4) even
16 the D/L increased from 0.021 to 0.032, compared to an increase by 22% (from 51.6 ft4 to 66.5
17 ft4) when the slab thickness increased from 10 in. to 12 in. with D/L equals to 0.021. This
18 confirms that the increase of slab thickness can enhance the stiffness of the composite girder
19 effectively leading to further reduction to the vibration of the structure.
20
21 CONCLUSIONS
22 This paper presents results of a comprehensive evaluation of the deflection criteria from various
23 Codes using a 3-D dynamic model. A parametric study was conducted to identify the most
24 sensitive parameters to control the vibration of steel girder bridges. Based on the analysis results,
25 the following conclusions could be made:
26 Neither depth-to-span limitations nor static deflection criteria could provide an effective
27 vibration control of steel girder bridge under normal truck traffic conditions.
28 If the depth-to-span ratio limit is not considered, the more economic design can be
29 achieved by choosing shallower HPS girder sections while controlling the bridge
30 vibration well under the perceptible level by increasing the concrete slab thickness.
31 Bridges with the same first natural frequency may experience large variation in their
32 maximum vertical accelerations under different truck loads. Therefore, the first natural
33 frequency of the bridge, as used by the Canadian as well as Australian Codes, may not
34 effectively control bridge vibrations.
35 The vertical acceleration of HPS girder bridges can be effectively controlled by
36 increasing the mass and stiffness of the composite girder (i.e. increase concrete deck
37 thickness) without the need to increase HPS girder depth.
38
39 ACKNOWLEDGEMENTS
40 The research project related to instrumentation and monitoring of the Doremus Avenue Bridge is
41 sponsored by the New Jersey Department of Transportation (NJDOT) that is gratefully
42 acknowledged. The financial support and the technical assistance of NJDOT staff Harry Capers,

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 9

1 Jr. (retired), Nick Vittilo (retired), Jose Lopez (retired), W. Lad Szalaj (retired), and Dick Dunne
2 is gratefully acknowledged. The help of previous graduate students Nakin Suksawang, Joe Davis,
3 and Talat Abu-Amra, is also acknowledged. The analytical work by the second author was done
4 prior to his employment at the Bureau of Reclamation, the U.S. Department of Interior. The
5 findings expressed in this article are those of the authors and do not necessarily reflect the view
6 of NJDOT.
7
8 REFERENCES
9 1. AASHTO (2008). Load Resistance and Factor Design, Bridge Design Specifications.
10 America Association of State Highway and Transportation Official, Washington, D.C.
11 2. New Jersey Bridge Design Specifications, (2006), New Jersey Department of
12 Transportation, Trenton, NJ.
13 3. Demitz, J. R., Mertz, D. R., and Gillespie, J. W. (2003, March). Deflection Requirements
14 for Bridges Constructed with Advanced Composite Materials. Journal of Bridge
15 Engineering, 8(2), 73-83.
16 4. Nassif, H. H. (1993). Live Load Spectra for Girder Bridges. Ph.D. Dissertation,
17 Department of Civil Engineering, University of Michigan, Ann Arbor, Michigan, pp. 259.
18 5. Gindy, M. (2004). Development of a Reliability-Based Deflection Limit State for Steel
19 Girder Bridges. Ph.D. Dissertation, Department of Civil Engineering, Rutgers, The State
20 University of New Jersey, Piscataway, New Jersey, pp. 267.
21 6. Reiher, H. and Meister, F. J. (1931). The Effect of Vibration on People. (in German(,
22 Forschung auf dem Gebeite des Ingenieurwesens, Vol. 2, No. II, P381. Translation:
23 Report No. F-TS-616-RE, Headquarters Air Material Command, Wright Field, Ohio,
24 1946.
25 7. Goldman, D. E. (1948). A Review of Subjective Responses to Vibratory Motion of the
26 Human Body in the Frequency Range 1 to 70 Cycles per Second. Naval medical research
27 institute, National naval medical center, Bethesda, MD.
28 8. Janeway, R. N. (1950). Vehicle Vibration Limits for Passenger Comfort. From Ride and
29 Vibration Data, Special Publications Department (SP-6), Society of Automotive Engineers,
30 Inc., p. 23.
31 9. Oehler, L. T. (1957). Vibration Susceptibilities of Various Highway Bridge Types.
32 Michigan State Highway Department (Project 55 F-40 No. 272).
33 10. Oehler, L. T. (1970, February). Bridge Vibration Summary of Questionnaire to State
34 Highway Departments. Highway Research Circular. Highway Research Board (No. 107).
35 11. Wright, D. T. and Green, R. (1964, May). Highway Bridge Vibration. Part II: Report No.
36 5 Ontario Test Programme. Ontario Department of Highways and Queens University.
37 Kingston, Ontario.
38 12. Wright, R. N. and Walker, W. H. (1971, November). Criteria for the Deflection of Steel
39 Bridges. Bulletin for the America Iron and Steel Institute, No. 19.

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 10

1 13. Ontario Ministry of Transportation, Quality and Standards Division (1995) Ontario
2 Highway Bridge Design Code/Commentary and March 1995 update, Quality and
3 Standards Division, Toronto, Ontario, Canada.
4 14. Nassif, H., Suksawang, N., Davis, J., Gindy, M., Abu-Amra, T. (2010) Instrumentation,
5 Field Testing, and Monitoring of the Doremus Avenue Bridge: Superstructure, FHWA-
6 NJ-2005-13, Final Report Submitted to NJDOT, 155 pp.
7 15. American Society of Civil Engineering. (1958, May). Deflection Limitation of a Bridge.
8 Journal of the Structural Division, 84, (Rep. No. ST3).
9 16. Wu, H. (2003) Influence of Live-Load Deflections on Superstructure Performance of Slab
10 on Steel Stringer Bridges. Ph.D. Dissertation, Department of Civil and Environmental
11 Engineering, West Virginia University
12 17. Nowak, A. S. and Grouni, H. N. (1988). Serviceability Considerations for Guideways and
13 Bridges. Canadian Journal of Civil Engineering, 15(4), 534-537.
14 18. Ministry of Transportation, Quality and Standards Division (1991) Ontario Highway
15 Bridge Design Code/Commentary, (3rd edition). Toronto, Ontario, Canada.
16 19. CSA International (2000, May). CAN/CSA S6- 00 and Commentary. Canadian
17 Highway Bridge Design Code. Canadian Standards Association, Toronto, Ontario,
18 Canada.
19 20. 92 AUSTROADS BRIDGE Design Code (1992). SECTION TWO-CODE Design Loads
20 and its COMMENTARY, AUSTROADS, HAYMARKET, NSW, AUSTRALIA.
21 21. 96 AUSTRALIAN BRIDGE Design Code (1996). SECTION SIX-CODE Steel and
22 Composite Construction, AUSTROADS, HAYMARKET, NSW, AUSTRALIA.
23 22. Oriard, L. L. (1972). Blasting Operations in the Urban Environment. Bulletin of the
24 Association of Engineering Geologists, IX (1.), pp 27-46
25 23. Burke, M. P. (2001). Superstructure Flexibility and Disinetegration of Reinforced
26 Concrete Deck Slabs: An LRFD Perspective, Transportation Research Record No. 1770,
27 pp. 76-83, Paper No. 01-0132.
28 24. Roeder, C. W., Barth, K. B., and Bergman, A. (2002, May). Improved Live Load
29 Deflection Criteria for Steel Bridges. Final Report NCHRP 20-07/133, University of
30 Washington, Seattle, WA.
31 25. Azizinamini, A., Barth, K., Dexter, R., Rubeiz, C. (2004). High Performance Steel:
32 Research Front- Historical Account of Research Activities. Journal of bridge engineering,
33 Vol. 9, No.3, p212-217.
34 26. Nassif, H. H., Liu, M., and Ertekin, O. (2003, March). Model Validation for Bridge-
35 Road-Vehicle Dynamic Interaction System. ASCE Journal of bridge engineering, Vol. 8,
36 No. 2, p112-120.
37 27. Nassif, H. and Nowak, A. (1995) Dynamic Load Spectra for Girder Bridges,
38 Transportation Research Board 1476, TRB, National Research Council, Washington, D.
39 C., pp. 69-83.
40

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 11

1 LIST OF FIGURES
2 FIGURE 1 Cross-Section of Instrumented Span Of The Doremus Avenue Bridge
3 FIGURE 2 Static Deflection vs. First Flexural Frequency (Ministry, 1991; CSA International,
4 2000; AUSTROADS, 1992; AUSTRALIAN, 1996) [18, 19, 20, 21] (U=Unacceptable,
5 A=Acceptable)
6 FIGURE 3 Validation of 3-D Dynamic Model by Measured Stresses at Span 2, Girder 9
7 FIGURE 4 Measured Acceleration Records on Span 3, Girder 8 (Interior), Middle Span (Gindy
8 2004) [5]
9 FIGURE 5 Configurations of Truck A and B
10 FIGURE 6 Maximum Static Deflection vs. Depth-to-Span Ratio (D/L)
11 FIGURE 7 Maximum Vertical Acceleration vs. Depth-to-Span Ratio (D/L)
12 FIGURE 8 Acceleration Records with Identical Slab Thickness
13 FIGURE 9 Acceleration Records with Different Slab Thickness
14 FIGURE 10 Maximum Static Deflection vs. First Natural Frequency
15 FIGURE 11 Maximum Vertical Acceleration vs. First Natural Frequency
16 FIGURE 12 Variation of Maximum Vertical Acceleration vs. Girder Moment of Inertia.
17

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 12

220
CSA-w/o sidewalk
200
180 CSA-with sidewalk, little pedestrian use
Static Deflection (mm)

160 CSA-with sidewalk, significant pedestrian use


140
AUSTRALIAN
120
100
80 U
60 U
40 A
20
A
0
0 2 4 6 8 10

1
First Flexural Frequency (Hz)
2 Figure 1. Static Deflection vs. First Flexural Frequency (Ministry, 1991; CSA International,
3 2000; AUSTROADS, 1992; AUSTRALIAN, 1996) [18, 19, 20, 21] (U=Unacceptable,
4 A=Acceptable)
5

6
7
8 Figure 2. Cross-Section of Instrumented Span of the Doremus Avenue Bridge

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 13

1
2 Figure 3. Validation of 3-D Bridge Dynamic Model using Measured Stresses at Span 2, Girder 9
3

First natural frequency = 1.59


H

4
5 Figure 4. Measured Acceleration Records on Span 3, Girder 8 (Interior), Middle Span (Gindy
6 2004) [5]

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 14

1 2 3 4 5
14-8 4-7 23-4 4-6 Truck A

10.76 kips A 15.74B15.74 C 17.05D17.05


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
1
2 Figure 5. Configurations of Truck A and B
3
1.2 0.030
t = 8.66"

Maximum Static Deflection (m)


t = 10"
Maximum Static Deflection (in.)

1.1
D / L < 0.027
t = 12"
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)
4
5 Figure 6. Maximum Static Deflection vs. Depth-to-Span Ratio (D/L) (t = Concrete Slab
6 Thickness)
7

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 15

70

Maximum Vertical Acceleration (in./sec.2)

Maximum Vertical Acceleration (m/sec.2)


t = 8.66" 1.7
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)


1
2 Figure 7. Maximum Vertical Acceleration vs. Depth-to-Span Ratio (D/L) (t = Concrete Slab
3 Thickness)
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.)
4
5 Figure 8. Acceleration Records with Identical Slab Thickness ((t = Concrete Slab
6 Thickness))

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 16

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

Acceleration (m/sec.2)
t = 12" 0.4
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.)
1
2 Figure 9. Acceleration Records with Different Slab Thickness (t = Concrete Slab
3 Thickness)
1.2 0.030
t = 8.66"
Australian

Maximum Static Deflection (m)


Maximum Static Deflection (in.)

1.1 t = 10"
bridge design
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 Frenquency (Hz)
4
5 Figure 10. Maximum Static Deflection vs. First Natural Frequency (t = Concrete Slab
6 Thickness)
7

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 17

70

Maximum Vertical Acceleration (in/sec.2)

Maximum Vertical Acceleration (m/sec.2)


Truck A 1.7

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 Frenquency (Hz)
1
2 (a) Concrete Slab Thickness: 8.66
Maximum Vertical Acceleration (in/sec.2)

70

Maximum Vertical Acceleration (m/sec.2)


Truck A 1.7

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
3 First Natural Frenquency (Hz)
4 (b) Concrete Slab Thickness: 10

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 18

Maximum Vertical Acceleration (in/sec.2)

Maximum Vertical Acceleration (m/sec.2)


70
Truck A 1.7

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
1 First Natural Frenquency (Hz)
2 (c) Concrete Slab Thickness: 12
3 * Based on 0.5 (Frequency)
4 Figure 11. Maximum Vertical Acceleration vs. First Natural Frequency

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 19

Moment of Inertia (m4)


0.17 0.27 0.37 0.47 0.57 0.67
70
Maximum Vertical Acceleration (in./sec.2)

Maximum Vertical Acceleration (m/sec.2)


t = 8.66" 1.7
Unpleasant
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

1 Moment of Inertia (ft4)


2 Figure 12. Variation of Maximum Vertical Acceleration vs. Composite Girder Moment of
3 Inertia (t = Concrete Slab Thickness)
4
5

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 20

1 LIST OF TABLES
2 TABLE 1 Maximum Vertical Acceleration for Different Depth-To-Span Ratios And Slab
3 Thickness Under Various Truck Loading Conditions
4

TRB 2011 Annual Meeting Paper revised from original submittal.


Nassif, Liu, Su, and Gindy 21

1 Table 1. Maximum vertical acceleration for different depth-to-span ratios and slab thickness
2 under various truck loading conditions
Steel Girder Depth, m (in.) 1.44 (56.69) 1.20 (47.24) 0.96 (37.79) 0.72 (28.34)
D/L= 0.032 0.027 0.021 0.016
Satisfy minimum D / L =
Yes Yes No No
0.027?

Concrete Slab Thickness


Maximum Acceleration, m / sec.^2 (in./sec.^2)
8.66 in (0.220 m)
Truck A 1.39 (54.79) 1.49 (58.64) 1.37 (53.82) 1.42 (55.75)
Truck B 0.91 (35.79) 1.05 (41.23) 0.94 (37.02) 1.46 (57.42)
Trucks A & B (Side by Side) 1.51 (59.60) 1.52 (59.91) 1.40 (55.21) 1.41 (55.53)

Concrete Slab Thickness 10


Maximum Acceleration, m / sec.^2 (in./sec.^2)
in (0.254 m)
Truck A 1.00 (39.30) 1.01 (39.91) 1.02 (40.29) 1.16 (45.80)
Truck B 0.77 (30.13) 1.05 (41.46) 0.89 (34.99) 0.84 (33.26)
Trucks A & B (Side by Side) 1.14 (45.05) 1.15 (45.08) 1.18 (46.47) 1.18 (46.36)

Concrete Slab Thickness 12


Maximum Acceleration, m / sec.^2 (in./sec.^2)
in (0.305 m)
Truck A 0.71 (28.09) 0.73 (28.84) 0.80 (31.47) 0.86 (33.74)
Truck B 0.73 (28.86) 0.69 (27.23) 0.76 (29.89) 0.63 (24.80)
Trucks A & B (Side by Side) 0.71 (28.14) 0.73 (28.55) 0.73 (28.89) 0.77 (30.24)
3
4
5

TRB 2011 Annual Meeting Paper revised from original submittal.

Vous aimerez peut-être aussi