Vous êtes sur la page 1sur 8

Construction and Building Materials, Vol. 12, Nos 2 3, pp.

. 143 150, 1998 1998 Elsevier Science Ltd Printed in Great Britain. All rights reserved 0950 0618 r 98 $19.00 q 0.00

PII:S09500618(97)000160

Fatigue and dynamic load measurements on modular expansion joints


C. W. Roeder

Department of Civil Engineering, University of Washington, 233 More Hall, FX-10, Seattle, WA 98195, USA Received 20 May 1994; accepted 5 July 1997
Modular expansion joints are commonly used on bridges with expansion and contraction movements in excess of 127 mm. These joints ensure a serviceable bridge surface, and protect the structure and substructure against water damage. However, these joints are dynamically loaded by truck wheels crossing the joint, and severe fatigue damage has been noted. The various types of modular joints are discussed, and fatigue damage on one joint is described. Analytical studies considering the static and dynamic behaviour of the joint are summarized, and the results are correlated to fatigue design methods. Preliminary results from an experimental study of field measurements of joint behavior are provided. The results indicate that the load spectrum is a very important element in the fatigue life of these joints. Acceleration and braking of the truck traffic induces large horizontal forces on the joint, and is a primary contributor to the fatigue damage. 1998 Elsevier Science Ltd. All rights reserved. Keywords: bridges; dynamic response; expansion joints

Introduction
Modular expansion joints are often used on bridges whose expected movements are larger than 127 mm. These systems are relatively expensive, since they are sealed against moisture penetration and are believed to reduce corrosion associated with water penetrating into the structure and substructure. They are separated from other sealed joint systems because they are primarily a structural system designed to accommodate large movements and support large dynamic wheel loads. That is, the design of these joint systems requires consideration of a range of structural design factors including dynamic wheel loads, stress and strain in components, and dynamic response of the joint rather than intuitive rules or guidelines. Modular expansion joints have been used on bridges with movements in excess of one meter, but fatigue cracking has been observed on many of these joints1 3 , and the cracking has signicantly reduced the fatigue life of the expansion joint system. This paper will provide an overview of the fatigue damage noted on one particular bridge, and a summary of the analysis performed to examine the fatigue cracking. Various fatigue design methods for modular expansion joints will be discussed, and a brief description of an experimental program to examine the fatigue loading characteristics will be provided. 143

Modular expansion joints


There are three major types of modular expansion joints, but all three types have some common features. First, the riding surface is accommodated by centerbeams which are transverse to the direction of the roadway, and are closely spaced as illustrated in the photo of Figure 1. The centerbeams are usually heavy rail sections, which move further apart or closer together as needed to accommodate expansion or contraction of the bridge. A rubber membrane is placed between each centerbeam to seal the joint system. Three different types of support mechanisms are used to support the centerbeams while accommodating the required movements as illustrated in Figure 2. Mechanical systems such as the scissors-like system illustrated in Figure 2(a) have been built to expand and contract while accommodating movements and providing very stiff gravity load support to the centerbeams. The scissors-like action of the support mechanism maintain the horizontal position and spacing of the centerbeams, while accommodating the required movements. The other two systems illustrated in Figure 2(b) and (c) use a support bar as a transverse beam to support the centerbeams. The support bar is a stiff, strong rectangular steel section. In some cases, the support

144

Fatigue and dynamic load measurements on modular expansion joints: C. W. Roeder


vehicle loading. Multiple support bar systems have separate support bar for each support point for each centerbeam. The centerbeam is rigidly connected to the support bar with multiple support bar systems. Single support bar systems have a single support bar which supports all centerbeams at each support point. The centerbeams must have a moveable attachment with stirrups, elastomeric springs and low friction sliders between the centerbeam and support bar. Most modular expansion joints that are used in the United States today employ a support bar system. The size of the centerbeam and the support bar depend upon the dynamic loadings of vehicles crossing over the joint and the span or expansion distance of a joint. Normal design stress checks must be made to size the members, and fatigue is an important consideration in the modular joint design. Multiple support bar systems have rigid attachments between the centerbeam and the support bar since the support bar is dedicated to a particular centerbeam. This connection and adjacent areas of the centerbeams and support bars have been locations of fatigue cracks in some bridges. Single support bar systems have had fatigue cracking in centerbeams in and around the stirrup connections as illustrated in the photo of Figure 3. Unfortunately, the fatigue design criteria for these joints has not been

Figure 1 Photo of the road surface of a modular expansion joint

bar is pinned at one end to the bridge superstructure, and on a low friction sliding surface at the other end. In other cases, the support bar is on a low friction sliding surface at both ends, and springs and damping devices are used to control the horizontal position of the support bar and the centerbeam spacing. The bridge movements are accommodated through this support bar sliding action. The support bars are spaced at fairly close intervals commonly 1.5 m.. This spacing limits the spans and bending stress of the centerbeams under

Figure 2 Three types of modular joint: a. mechanical support system; b. multiple support bar system; and c. single support bar system

Fatigue and dynamic load measurements on modular expansion joints: C. W. Roeder

145

Figure 3 Photo of fatigue cracking

fully established yet. The fatigue problem is quite different for modular joints than its normal steel bridge fatigue design since many, many more cycles of loading will occur, and the fatigue loading depends on the dynamic wheel load rather than the truck weight. The 3rd Lake Washington Bridge in Seattle, Washington, has two 1200-mm modular expansion joints as shown Figure 1 at opposite ends of 1.75 km of oating pontoons. The bridge was opened to trafc in June 1989, and these joints utilize the single support bar swivel joist design. Steel tubes were substituted for the I-shaped centerbeams used in the original design because domestically produced centerbeams were unavailable and FHWAs Buy American steel requirements for federally funded bridge construction would not permit the use of foreign steel. Approximately 6 months after the bridge was opened to trafc, there were numerous complaints of expansion joint noise. Inspection of the joints showed that some elastomeric bearings used to cushion the trafc impact between the centerbeams, stirrups and support bars were loose. Shims were added, but within 1 year cracks in the tubular centerbeams were observed. Most of these cracks started at the toe of the stirrup to centerbeam llet weld and progressed through the centerbeam, as shown in Figure 3. One crack occurred at the end of a ange cover plate. The manufacturer repaired seven of these cracks in April 1991 by rewelding the cracked metal. Additional cracks were noted in the centerbeams after this rst repair, and seven more cracks were repaired in November 1991. Additional cracks were noted after this second repair, and some of the previously repaired cracks reappeared. In the intervening period, more than 20 fatigue cracks have been noted. This research was initiated in response to this fatigue cracking.

including impact proposed by Tschemmernegg are: a vertical downward load of q 91.0 kN, a minimum vertical rebound load of y 27.3 kN and a horizontal load of q 18.2 kN. These design loads for fatigue are based on eld measurements from several bridges in Europe 6 8. The stresses then are calculated at critical locations to determine the maximum computed stress range, max . The center beams are treated as continuous beams, and the elastomeric springs and bearings are treated as rigid supports for determination of the moments and the stress level. For normal conditions, each centerbeam carries approx. 50 60% of the wheel load with a 1.8 m wheel spacing since the wheel distributes the load to more than one centerbeam. Note that the primary loading considered in the design method produces compressive stress in the same area as the fatigue cracking on the 3rd Lake Washington Bridge. Curves representing stress-range to number of cycles S N curves. are determined for each critical component or location. The S N curve is constructed with a slope of y 0.33 on the log log S N curve for a stress range of less than 5 million cycles and a slope of y 0.20 for a stress range between 5 million and 100 million cycles. The intercept at 100 million cycles is the theoretical endurance limit, L , proposed by Tschemmernegg 4 6 , but all tests are performed at 2 million cycles or less. The maximum calculated stress range is compared to the theoretical endurance level from the S N curves by max - L 2 1.

Review of previous work


A relatively simple fatigue limit states design method has been proposed by Tschemmernegg 4 6 . First, the loads on the bridge and the expansion joint are determined. The design limit states fatigue wheel loads,

where max is the calculated stress range based on the dened range of wheel loads, and L is the limit states fatigue stress range at 100 million cycles. The maximum calculated stress range is divided by two because of the partial safety factors Mf and , where accounts for accumulated fatigue damage. There are reasons for questioning the validity of the Tschemmernegg procedure. Koster 9 believes that the elastic deformation of the system affects the stress distribution and the fatigue potential. He contends that deformability of the joint is desirable because it may spread the load and possibly reduce the critical fatigue stress. However, Tschemmernegg contends that the elastic deformation and stress distribution will not occur because of the very short duration of the wheel loads. Agarwal performed a series of eld measurements on a modular expansion joint on a bridge in Ontario, Canada10 . These eld measurements suggested that the loads and load spectrum recommended by Tschemmernegg may not be universally applicable. Large horizontal forces noted by Tschemmernegg 6 were not detected and the load range and spectrum were different. However, the centerbeam instrumentation that Agarwal used may not have been adequately located or sensitive enough to detect horizontal loads on the joint.

146

Fatigue and dynamic load measurements on modular expansion joints: C. W. Roeder


Local nite element analyses The global analyses did not provide a complete picture of the state of stress in the critical stirrup location. The centerbeam and the stirrup region were modeled with a detailed local model. The centerbeam was modeled with shell elements, and the stirrups were modeled with three-dimensional brick elements. The loads at the ends of the tube and the loads attributable to the elastomeric springs were obtained from the global computer analysis results. The local analysis showed considerable local bending deformation of the walls of the tube near the stirrup weld. The bending stresses caused by these plate bending moments were found to be approximately the same magnitude as the basic beam bending stress described earlier in the global analysis. These bending stresses varied from tension to compression through the thickness of the wall of the tube and caused increasing stress tensile . on the inside of the tube and decreasing stress on the outside of the tube near the stirrup weld. The tubular centerbeams contribute to the fatigue problem because of local deformation and through-thickness plate bending stress, but fatigue would have been a problem even if another section had been used for the centerbeams because of the centerbeam span length. Dynamic analysis The global model was used to perform dynamic analysis on the modular joint. The mass of the components of the modular expansion joint were added to the model and many modes of vibration were computed. The dynamic modal computations required a large amount of computer time because of the broad distribution of mass and stiffness and the large number of degrees of freedom. The modes of vibration for this modular expansion joint were closely spaced with hundreds needed to include the predominate portion of the mass in three-dimensional vibration. The longest periods were associated with horizontal movement. In examining the dynamic response of the joint, it was noted that the wheel load on any centerbeam is initially zero until the wheel makes contact with the given beam. The load reaches its maximum value when the wheel is nearly centered over the given centerbeam, and it then decreases until the wheel separates from the beam. If the truck is traveling at a constant velocity, this translates into the linear time dependent ramp load function shown as an insert in Figure 5. For a tire contact length of 24 cm and a centerbeam width of 80 mm, vehicles at 33, 67 and 100 kmrh would have load durations of 0.035, 0.017 and 0.012 s, respectively. The gure shows the maximum dynamic response divided by the static response as a function of the ratio of the duration of the ramp function loading to the period of the system. The dynamic amplication of 1.0 indicates that the structure feels the full static loading, and a factor greater than 1.0 implies impact or dynamic amplication. The centerbeam feels the full static load and potential impact if the duration of the loading is

Figure 4 Global computer model for nite element analysis

Analysis of the modular joint


Static and dynamic computer analyses of the large modular joint on the 3rd Lake Washington Bridge were performed. The analyses included the global behavior and the local stress and strains in the system. A global model as illustrated in Figure 4 was analyzed rst. The centerbeams, support bars, and stirrups were modeled with beam elements. The geometry and member properties were based on information obtained from the contract shop drawings. The elastomeric bearings were modeled as compression and shear springs, where the spring stiffness was determined by typical models of elastomeric bearing stiffness 11 . The drawings did not specify the elastomer properties, and a study was performed to determine the sensitivity of the computed response to variations in the elastomer properties. The elastomeric spring stiffness did not have a dramatic effect on the bending moments, but some aspects of the behavior noted with elastomeric springs was different from that noted when rigid connections between the support bars and centerbeams were employed. The bending moment and nominal stress range in the centerbeam were smaller when the load was placed on interior centerbeams than on edge centerbeams, because edge beams had alternating long and short spans and interior beams had more uniform span lengths. The more uniform spans reduced the range of the stress and the bending moments at the critical stirrup locations. Thus, the edge centerbeams experience earlier fatigue cracking than the interior centerbeams. Global analyses were also performed with horizontal loads. Torsional deformation and weak axis bending of the centerbeams resulted when these horizontal loads were applied at the top of the center beam rail. Wheel loads cause multiple stress cycles for a single truck passage. Therefore the fatigue evaluation procedure of modular expansion joints must be different than that used for bridge girders. The analyses showed that a larger stress range is possible between different trucks because the trucks do not travel over the same path across the joint. This variability may double the stress range over that computed for a single wheel load.

Fatigue and dynamic load measurements on modular expansion joints: C. W. Roeder

147

Figure 5 Dynamic amplication of loading

longer than approximately 30% of the dynamic period of the structure. If the duration is less than 10% of the period, less than 30% of the static load is felt. Comparison of the computed natural periods with the load history and dynamic amplication illustrated in Figure 5 suggests that signicant amplication of horizontal forces should be expected at slower vehicle speeds. High-speed vehicles may cause the expansion joint to experience attenuation of horizontal loading. Amplication of the vertical response will occur over a wide range of vehicle speeds. It should be emphasized that these observations are meaningful for this particular expansion joint because of the transverse exibility of the system. Other modular joint systems, particularly multiple support bar joints, may be stiffer transversely and experience a greater dynamic amplication of horizontal loads.

Correlation of computed stress to fatigue criteria


The computed stress ranges were correlated with existing fatigue criteria. Normal AASHTO fatigue design is based on 2 million repetitions of the HS-20 truck loading. The stress range is the difference between the maximum stress due to the load and its impact and the unloaded condition. The welded stirrup to centerbeam

detail is somewhat analogous to detail 17 in the AASHTO Specications12 where attachments are welded to a longitudinally loaded member with short llet welds. Detail 17 indicates fatigue category D or E. Two million truck passes will cause far more than 2 million cycles of wheel loading. Comparison of the computed stress range with the HS-20 wheel loads produced stress ranges which were applied to the S-N curves for Categories C, D and E. The fatigue life which resulted was not consistent with the number of trucks which have crossed the bridge during the limited life service of the joint. This suggests that either the S N curves are not appropriate for this modular joint or the dynamic loads are quite different from the HS-20 load history postulated in the US design method. The Tschemmernegg fatigue design loads and the S N curve shown in Figure 6, were used to predict a fatigue life of 10 million cycles of total truck wheel loading for the as-built stirrup to centerbeam connection. This estimate is different from the AASHTO12 and AASHTO LRFD13 life estimates in that it includes the total number of truck passings and an estimate of accumulated damage. The accumulated damage estimate is based on a design wheel load spectrum proposed for expansion joints in Europe. Fatigue cracks were noted approximately 18 months after the bridge was opened to trafc, and 10 million cycles would require approximately 18 000 axles for one lane of trafc per day. A trafc count performed in 1990 found that the three westbound lanes of the bridge experienced approximately 6720 axles of bus and truck trafc during the busiest 12-h period of a normal work day. When the trafc was distributed over three lanes and the lighter weekend trafc was considered, the accumulated trafc was less than 20% of the proposed fatigue life estimate 14 . Furthermore, the cracks obtained in the laboratory fatigue test were quite different than those observed on the 3rd Lake Washington Bridge. This observation suggests that the Tschemmernegg procedure may not be applicable for these joints. While the Tschemmernegg method does not duplicate the fatigue cracking noted in the 3rd Lake Washington Bridge, the stress ranges predicted by the test may be fairly realistic. The reason for the approximate

Figure 6

S N curve developed for subject joint by Tschemmernegg method

148

Fatigue and dynamic load measurements on modular expansion joints: C. W. Roeder


will detect all trucks in the right hand lane. The disk drive is portable and it is carried into the University of Washington campus when it is full. The data is transferred again by an HPUX 700-340 computer, and can be analyzed by the normal research computer facilities.

accuracy of the S N curve is that the modular joint details are likely to always be close to AASHTO categories D or E because the weld detail is similar to AASHTO fatigue details 9 and 17. The detail may be closer to category D or even category C if the modular joint is less susceptible to fatigue and closer to E if it is more susceptible. However, it is clear that the load history is more important for establishing a fatigue design criteria and it may be quite different for US truck trafc. As a result, an experimental study was started to evaluate the wheel loads experienced by the modular joint under normal trafc conditions.

Test programs
Two types of tests were performed. First, two series of controlled tests were performed. During 17 19 August 1993, the right lane of trafc was shut down for several hours each day, and loads were applied by a moderately heavy, three-axle dump truck. The dimensions and the static wheel loads of the truck were measured before testing. The truck passed over the joint at known speeds and locations. The truck was sometimes at constant speed, braking, accelerating, or at rest. A second series of controlled tests was performed during 1 2 February 1994. This second series of controlled tests were performed when the bridge was closed to other trafc. Forty-two load passes were made with the same truck used in the earlier tests and a nearly identical loading. Nearly all of the truck passes were made at various points within the outside southern most. lane, so that the maximum useful information with regard to truck location could be obtained. Most of the tests were performed with eastbound truck trafc, but a few passes were performed with a westbound truck. These tests may be useful in interpreting future reversible lane measurements. Two tests were performed with the truck passing in the center lane so that the effect of such a truck passing on the measured results could be determined. The results of these tests will be used to establish basic elements of joint behavior such as the effect of truck position, truck braking or acceleration, and distribution of load between centerbeams. The results will also be used as an aid in interpreting the second type of the test results. The second type of tests consists of measurements of the uncontrolled truck and bus trafc and its affect on the joint. The recorders were set to record data when a large truck crossed the joint in the outside lane. The signal was set such that a truck in the middle lane would not trigger the device nor would a heavy car or light truck in the right lane. Data was recorded continuously in blocks of 10 trucks. A few trucks may be missed during the short period required to transfer the block of data from the wave form recorder to the computer, but the percentage is small. Furthermore, a few lighter trucks are missed because they are not heavy enough to trigger the instruments, but they do not contribute much to fatigue. To illustrate the number of trucks recorded, it can be noted that during the 24-h period on Wednesday, 1 September, 990 vehicles were measured in this right hand lane of the eastbound I-90. This is not inconsistent with the truck trafc count of 13 November 1990 where a total of 6720 axles were counted in all of the westbound lanes during a 12-h period. The 990 trucks should result in nearly

Experimental program and measurements


The large modular expansion joint at the east end of the eastbound lanes of the I-90 3rd Lake Washington Floating bridge were instrumented during the Summer 1993. The general objectives of the measurements are to 1. verify the results of the computer analysis; 2. determine the load spectrum for vertical and horizontal loads for fatigue design and evaluation of modular expansion joints; 3. determine the stresses and strains of critical joint components under the applied loads; and 4. verify the dynamic characteristics of the modular joint system including impact and damping. Two types of instrumentation were used on this joint and two types of tests or measurements are being performed. The rst type of instrumentation includes 17 groups of four strain-gages which are connected as full Wheatstone bridges. Most of these channels measure bending in the vertical load plane, but two measure bending in the horizontal load plane. Three other channels measure local effects around the stirrup to centerbeam connection. The second type of instrumentation include two LVDTs which were used to measure horizontal displacement of the centerbeams. HP5813A wave form recorders were used to record most of the data and additional data was recorded with a HP3852A data acquisition system. The wave form recorders are capable of recording up to 4 000 000 samples per second of data per channel, but they are sampling at the rate of 2000 samples per second for these tests since this is more than ample to measure the joint response. The recorders are coupled together and they are self triggering. That is, they can be continually sampling data but they only record data when a big enough measurement is noted. At this time a short burst at present 2 s. of data is recorded. The data recorders can record a number of trucks until their internal memory is full. The data is then transferred to an HP9816 computer and it is stored in a compact binary format in an IEM 5300 disk drive. This operation is automatic and no intervention by the researchers is required. It can operate day and night and

Fatigue and dynamic load measurements on modular expansion joints: C. W. Roeder

149

Figure 7 Measured bending moment in centerbeam due to vertical wheel loads at 90 kmrh and static conditions

Figure 9 Measured bending moment in centerbeam due to horizontal wheel loads at 90 kmrh with moderate braking

5000 axles in a single lane for a 24-h period. These measurements will be used to establish a fatigue load spectrum for vertical and horizontal wheel loads on the joint. It should be emphasized that these loads are not static wheel loads. They are dynamic loads including the effect of impact and dynamic amplication or attenuation. The variation and distribution of stress is quite interesting since it affects fatigue behavior. However, the stress and bending moment at the strain gage locations depends upon the placement of the wheels on the centerbeam, and this placement is not known for the uncontrolled trafc measurements. The rst four channels of measurements allow determination of the vertical plane bending moment at four locations on the centerbeam under one lane of trafc. These four moments are enough to estimate both the magnitude of the dynamic wheel load on the centerbeam and position of the wheel on the joint through inuence lines developed for each gage location. It should be noted that the inuence lines are quite complicated and depend upon dynamic effects as well as static loading. More measurements are made than are actually needed to determine load and position. This provides redundancy which can be used to verify and improve the

accuracy of the estimates. The position of the vehicle wheels combined with the two horizontal plane bending moments allow estimation of the horizontal wheel load on the centerbeam.

Preliminary interpretation of the results


It must be noted that the analysis and interpretation of the results will continue for another year. Thus, any discussion of the results must be regarded as tentative and subject to change. However, some of the early test results have been examined. Figure 7 show the typical measured bending moments due the controlled test truck passing over the joint with nearly the same path very slowly and at 90 kmrh. The trucks were maintaining constant speed as they crossed the joint, and there was no braking or accelerating of the vehicle. The axle loads were approximately 9 tons for each of the two rear axles and 7 tons for the front axle. The dynamic load experienced by the joint is proportional to the maximum bending moment. Comparison of the two truck crossings show that there is 20 30% amplication of the vertical loads for the high-speed vehicle over the static loading. Figure 8 shows the horizontal bending moments due to the same two vehicles crossing the modular joint. Comparison of Figures 7 and 8 shows that the horizontal load is quite small 15% or less. compared to the vertical wheel load. The horizontal loads are much smaller than suggested by the Tschemmernegg fatigue design procedure. If the vehicle is braking or accelerating as it crosses the modular joint, the horizontal forces are much larger. Figure 9 shows the horizontal load bending moments with moderate braking with the truck at approximately 90 kmrh and Figure 10 shows the moments with emergency braking at a similar speed. Comparison of the gures shows that vehicle braking causes much larger horizontal dynamic loads than suggested by Tschemmernegg. No horizontal movement of centerbeam was noted if the truck was not braking or accelerating to gain speed over the joint. However,

Figure 8 Measured bending moment in centerbeam due to horizontal wheel loads at 90 kmrh and static conditions

150

Fatigue and dynamic load measurements on modular expansion joints: C. W. Roeder


for these joints will need to be quite different from that used for other fatigue design situations. The designer will need to consider many more cycles, and the loads and stress range depend upon local trafc patterns and the dynamic characteristics of the joint. The frequency of vehicle braking and acceleration across the joint becomes very important.

Acknowledgements
Figure 10 Measured bending moment in centerbeam due to horizontal wheel loads at 90 kmrh with emergency braking

This research is being funded by the Washington State Transportation Center TRAC. and the Washington State Department of Transportation. The assistance of John Van Lund, Myint M. Lwin and Alan H. Walley is gratefully acknowledged. The opinions and conclusions expressed or implied in the paper are those of the author and are not necessarily those of funding agency.

References
1 Roeder, C. W., Fatigue cracking in modular joints. Report No. WA-RD 306.1, Washington State Transportation Center TRAC., March 1993 2 Van Lund, J. A., Improving the quality and durability of bridge modular expansion joints. Transportation Research Record 1393, TRB, National Research Council, Washington, D.C., 1993, pp. 9 16 3 Mayrbaurl, R. M., Analysis of the Manhattan Bridge modular expansion joints. 1994 TRB Annual Meeting, Washington, D.C 4 Pattis, A. and Tschemmernegg, F., Fatigue testing and design of modular expansion joints. Report published by University of Innsbruck and The D.S. Brown Co., Innsbruck, Austria, March 1992 5 Fatigue design and testing for expansion joints. Technology Bulletin for Bridge Bearings, Expansion Joints and Components. No. 1, The D.S. Brown Co., North Baltimore, Ohio, October 1991, pp. 2 6 Tschemmernegg, F., The design of modular expansion joints. Preprint, Volume 1 Joints and Sealants, Third World Congress on Joint Sealing and Bearing Systems for Concrete Structures, Toronto, Canada, American Concrete Institute, October 1991, pp. 67 86 7 Ostermann, M., Stresses in elastically supported modular expansion joints under wheel impact load. Bauingenieur, 1991, 66, 381 389 in German. 8 Braun, C., Strain on expansion joints for road bridges under trafc loads. Bauingenieur, 1992, 67, 229 237 in German. 9 Koster, W., The principle of elasticity for expansion joints. In Joint Sealing and Bearing Systems for Concrete Structures, Vol. 2, SP-94. American Concrete Institute, Detroit, MI, 1986, pp. 675 711 10 Agarwal, A. C., Static and dynamic testing of a modular expansion joint in the Burlington Skyway. In Proceedings of the Third World Congress on Joint Sealing and Bearing Systems for Concrete Structures, Toronto, Canada, October 1991 11 Stanton, J. F., Roeder, C. W., Elastomeric bearings design, construction and materials. NCHRP report 248, TRB, National Research Council, Washington, D.C., 1982 12 Standard Specications for Highway Bridges, 15th ed. AASHTO, Washington, D.C., 1992 13 Proposed LRFD Standard Specications for Highway Bridges. 17 April 1992 Draft, NCHRP 12 33, Washington D.C., 1992 14 Pattis, A., Fatigue testing and design of modular expansion joints the BrownrMaurer swivel joist system repair procedures for 3rd Lake Washington Bridge. Short version of an unpublished report, August 1992

Figure 11 Measured horizontal displacement of centerbeam due to horizontal wheel loads at 90 kmrh with emergency braking

Figure 11 shows typical centerbeam movement if the truck is braking to an emergency stop. It can be seen that a large horizontal deection occurs under this severe braking condition. The maximum movement is approximately 10 mm and there is a permanent set of approximately 3 mm. The largest centerbeam movements appeared to occur at slower speeds because of the dynamic characteristics of the joint. This is consistent with the observations made in analysis that greater dynamic amplication of horizontal loads occurred at slower speeds because the duration of loading more closely matches the longer periods noted for horizontal displacement. The major portion of the deection is due to deformation and sliding of the elastomeric springs. That is, the centerbeam approximately moves as a rigid body. The elastomeric deection is elastic and is recovered after the load is removed, while sliding results in permanent set and is not recovered. This again is quite different than assumed in existing fatigue evaluations procedures. Additional interpretation of this data is needed. Ultimately, it is expected that a statistical load spectrum will be developed. Examination of the results to date clearly indicates that the fatigue evaluation procedures

Vous aimerez peut-être aussi