You are on page 1of 20

Web hosting

Custom Email

SiteBuilder

Integral Bridges
Disclaimer: Although the following recommended design procedure has been subjected to many rigorous tests and review, no warranty, expressed or implied, is made by the author as to the accuracy and functioning of the recommended design procedure nor shall the fact of posting on the web constitute any such warranty, and no responsibility is assumed by the author in any connection therewith.

What is an integral bridge?

Conventional Design Approach

Feasibility considerations

How to design an integral bridge


Consider the temperature variation and soil-structure interaction together !

Construction stages, loads and load combinations are important !

Analyse for the effects of temperature variation, earth and gravity loads

Analyse for the effect of seismic loads

Design Considerations
deck

abutment, wingwall and approach slab

abutment piles

bearings, pier and footing

Bibliography

Home

Curriculum Vitae

Engineering Sites

Universities

Employment Opportunities

Engineering Software

Useful Links

What is an integral bridge?


Integral bridges are single span or multiple-span bridges with a continuous deck and a movement system composed primarily of abutments supported on flexible piles. A typical two-span, prestressed-concrete-girder, integral bridge is shown below.

In these types of bridges, the road surfaces are continuous from one approach embankment to the other and the abutments are cast integral with the deck. The effect of forces parallel to the bridge longitudinal direction is minimized by designing the abutments and their foundations flexible and less resistant to longitudinal movements of the structure. Accordingly, the abutments are built shorter to reduce the restraint provided by the backfill soil to the longitudinal movement of the bridge. Only a single row of steel H piles is generally used to provide vertical support to abutments and minor resistance to longitudinal forces. The integral bridges described here are assumed to have slab-on-prestressed-concretegirder deck. The connection between the bridge deck and the abutment can be rigid or semi-rigid depending on the detailing of joint reinforcement. Elastomeric bearings are used

under each girder at intermediate supports. The reinforced concrete columns at intermediate supports may either be free standing or rigidly connected to a reinforced concrete cap-beam supporting the superstructure. The columns are assumed to be supported either by shallow foundations or deep foundations with two or more rows of piles.

Top of Page

Conventional Design Approach


In this section of the web site, the conventional design approach for integral bridges, currently used by many structural engineers in North America and Europe is briefly introduced. Its limitations attributed to the analytical models used for the analysis of the structure and design procedure are discussed. A typical three-span integral abutment bridge and its conventional analytical models are shown below.

The structure is analyzed only for the final stage assuming a completed structure.. The second sketch from the top of the above figure shows the continuous beam model used to analyze the structure for the design of the bridge deck only7,9-13. The piers and abutments in the actual structure are replaced with simple supports in the model. Thus, the continuity of the structure is totally neglected at pier and abutment locations for the deck design. A continuous frame model is only considered if specifically the design of deck-abutment continuity connections is required9-11,13. The weight of the concrete slab, girders, diaphragm beams and superimposed dead load and live load are considered for the design of the deck. The third sketch from the top in the above figure shows the conventional analytical model used for the design of deck-abutment joints assuming full composite action between the slab, girders and abutments. The analytical model does not consider a complete frame action since the piers are modeled as simple roller supports. The effect of continuity is considered only at the deck-abutment joints9,10,13. Furthermore, the model does not reflect the three dimensional effect of lateral loads on the piers, abutments, wing-walls and piles. A 3-D model is still necessary to analyze the structure for the effects of such loads. The weight of the slab, girders, diaphragm beams, superimposed dead loads, live load, earth pressure and effect of temperature variations are considered for the design of the deck-abutment joints. However, the correlation between the temperature variation and the magnitude of earth pressure is neglected in the design. The deck-abutment joints are designed by conservatively assuming a maximum passive earth pressure condition at the abutments. The detailing of abutment-deck continuity connections are standardized for a variety of range of applications14. Therefore, in most cases, the analytical model shown in the figure above for the deck-abutment joint design is not considered for the analysis of the structure. The structure is modeled as a simply supported continuous beam. The full design is completed based on that model. Nevertheless, the detailing may in fact vary depending on the type of loads applied to the structure. For example, seismic design provisions for reinforced concrete rigid frame structures usually require joint details that must have some level of rotational ductility for energy dissipation purposes15. Consequently, the standard reinforcement details for abutment-deck continuity connections may not be applicable if the structure is built in an area with high risk of seismic activity. The joint reinforcement details may also vary as a function of the structure geometry. In the design of piers supporting integral bridge decks, the unbalanced horizontal earth pressure forces resulting from unsymmetrical abutment configurations are assumed to be transferred directly to the approach embankments with no effect on piers. Accordingly, the piers are designed for vertical reaction loads transferred from the superstructure and for

lateral loads directly applied on the piers. Obviously, this design approach is limited to some simple cases where the structure is fully symmetrical and the soil pressures at both sides of the bridge are in equilibrium. The effect of axial load in the deck, resulting from earth pressure forces at both sides of the structure, is also neglected in the current design approach. The axial force applied to the bridge deck may cause extra shortening of the prestress concrete girders due to elastic deformation and creep. This may lead to a reduction in the effective prestressing force in the girders. The current design practice allows for the design of pile-abutment connection joints to develop full continuity. Consequently, bending moments at pile ends are produced due to temperature variations and vehicular traffic. These bending moments may be high enough to initiate plastic yielding of steel piles3,16-20. The repetitive variation of temperature and the effect of live load may cause low cycle fatigue in the piles 3. A hinge connection detail between the piles and the abutment may prevent such a potentially destructive problem. In the conventional design approach, the effect of seismic forces is usually neglected assuming that integral bridges are not prone to such forces because of their continuity9,10. However, earthquake excitations may cause remarkable rotations and settlements at the abutment foundations due to the flexible nature of single-row pile arrangement. Therefore, in some cases, a rigorous analysis may be required to assess the capacity of the structure to resist seismic forces.

Top of Page

Feasibility considerations
The foundation soil condition is an important factor in the feasibility study of integral bridges. The primary criteria of the decision-making process for integral bridge construction is the requirement of a single row of flexible piles to support the abutments. Accordingly, where the load bearing strata is not deep enough to allow piles longer than at least five meters in length, the site may not be considered suitable for integral bridges. If the soil is susceptible to slip failure, liquefaction, sloughing or boiling, integral bridges are not suitable for that site. In selecting the bridge types, consideration is given to the total length of the bridge, type of deck, type of traffic, location, and any unusual characteristics such as skew, curvature or grade. Integral bridges are generally suitable for total span lengths below 100 metres13. The limitation on length is mainly a function of the soil properties and seasonal temperature variations. It is imposed considering the ultimate resistance of abutments and piles to longitudinal movements and serviceability of the structure. Integral abutment bridges with skews greater than 35 degrees are not considered suitable for construction due to the nonuniform distribution of loads and difficulties in establishing the movements and their directions13.

Top of Page

How to design an integral bridge


Consider the temperature variation and soil-structure interaction together ! The earth pressure coefficient is a function of the displacement or rotation of the earth retaining structure. An integral bridge will experience elongation and contraction due to temperature variations during its service life. A very small displacement of the bridge away from the backfill soil can cause the development of active earth pressure conditions21,22. Therefore, when the bridge contracts due to a decrease in temperature, active earth pressure will be developed behind the abutment. When the bridge elongates due to an increase in temperature, the intensity of the earth pressure behind the abutment is a function of the magnitude of the bridge displacement towards the backfill soil. The actual earth pressure coefficient, K, may change between at rest, Ko and passive, KP, earth pressure coefficients depending on the amount of displacement. This earth pressure coefficient is expressed as a function of temperature varioation and soil-structure interaction as follows:

Where;

slope of the earth pressure variation. Its value varies as a function of the backfill soil type. Typical values for various soil types are provided elsewhere22,23. Ld span length of the bridge. we effective width of slab, in this case assumed to be equal to the spacing of girders. Eg modulus of elasticity of the girder material Ag cross-sectional area of the girder, As cross-sectional area of the portion of the deck slab with an effective width equal to the spacing of girders n modular ratio defined as the ratio of the elastic modulus of slab material to that of girder material. h abutment height, s unit weight of backfill soil coefficient of thermal expansion for the deck material T differential temperature variation

The actual earth pressure coefficient is calculated using the above equation. Since the equation yields a smaller earth pressure coefficient than passive, more economical designs for abutment and piles may be obtained.
Top of Page

Construction stages, loads and load combinations are important ! The construction of an integral bridge is done in stages. Therefore, it must be analyzed for each construction stage to ensure that the structure has adequate capacity to sustain the applied loads particular to the stage under consideration. Two stages are considered for the design of a slab-on-prestressed-concrete-girder integral bridge. The loads applied at each stage are listed in Table below. Summary of stage loading

Loads Applied in the Stage Stage # Stage Name Load ID Description

1 Simply supported beams

Own weight of girder

Pre-tensioning

3 Composite structure 2 4

Weight of wet concrete slab, diaphragms and ab

Superimposed dead load (e.g. in sidewalks, curbs,

Asphalt weight

Long term prestress losses

Highway live loading and sidewalk load, or pedestr Fatigue Limit State

As load 7 but at Serviceability Limit Stat

As load 7 but at Ultimate Limit State

10

Thermal load due to longitudinal expansio

11

Thermal load due to longitudinal contracti

12

Passive earth pressure

13

At rest earth pressure

14

Active earth pressure

15

Seismic loads

In the first stage the slab concrete is assumed to be wet. Accordingly, the prestressedconcrete girders alone resist the applied loads in this stage. The structure is analyzed for the effects of prestressing force, dead weight of the girders, weight of wet concrete slab, and

weight of the diaphragms. In the second stage the bridge is assumed to be in service. Full composite action is considered between the slab, girders and abutments. The effects of superimposed dead loads, asphalt weight, temperature variation, soil pressure and live load are considered in this stage. The most critical load combinations resulting in maximum responses at various locations on the structure in stage 2 are illustrated in the following figure. In the figure, the locations indicated by a solid circle are for flexural responses and the ones indicated by a short line are for shear responses.

The most critical load combinations are listed in the table below. In the table, the cells marked by M indicate the application of the specified load to obtain the optimum flexural response for the location under consideration. Similarly, the cells marked by V indicate the application of the specified load to obtain the optimum shear response for the location under consideration. It is noteworthy that the earth pressure loads applied at abutments are considered in correlation with thermal loads. At rest earth pressure is considered when there is no thermal movement but, passive and active earth pressures are considered when there is thermal expansion and contraction respectively. At some response locations along the structure (e.g. 1 & 4), the effect of temperature variation and earth pressure oppose one another. Therefore, all combinations of temperature and earth pressure effects must be considered to obtain the optimum response.

Load combinations for maximum flexural (M) and shear (V) responses

Response Location

Dead Loads

Live Loads

Temperature

Earth Pressur

Span #

Pos.

Zero

Neg.

Pas.

At Rest

M V

M V

M V

M V

M V

M V

M V

M V

M V

M V

M V

M V

M V

M V

M V V

M V

V M M

M V

M V V

M V

M V

Top of Page

Analyse for the effects of temperature variation, earth and gravity loads For the analysis of an integral bridge subjected to temperature variation, gravitational and earth pressure loads, a separate analytical model is considered for each construction stage. The following Figure illustrates a typical two-span integral bridge and its analytical model for construction stage one.

For this stage, the integral bridge is idealized as a 2-D structure considering only one girder. The naked girder alone is considered assuming that the deck concrete which provides continuity between the structural components is not hardened. The bridge is analyzed considering each span as a simply supported beam for stage one loads. The resulting internal element forces are then stored for superimposing them to the ones resulting from the loads applied in stage two. The following figure illustrates the same bridge and its analytical model for construction stage two.

The bridge is idealized as a 2-D structure considering only one girder and an effective width of slab. Accordingly, the abutments are idealized to have a tributary width equal to that of the slab. Similarly, the number of columns and piles per tributary width is calculated and their stiffness is lumped to obtain a single column or pile element for

analysis purposes. Full continuity at the intermediate supports and at the abutment-deck connection joints is considered assuming that the concrete is fully hardened. The idealized abutment and pier members are connected to the deck nodes by abutment-deck or pierdeck connection elements. The connection elements are used to define the rotational and/or translational stiffness of joints at various parts of an integral bridge. Normally, if adequate continuity is provided between the slab, girders and abutment using a proper reinforcement detailing, connection elements at abutment-deck joints are assumed as rigid. Based on the type of reinforcement detailing at the deck-abutment joint, adequate continuity may not be provided between the connected elements. In this case, the joint may be idealized as a hinge or semi-rigid by adjusting the stiffness of the connection element. The connection element at the deck-pier joint represents the bearing and the one at the pier base is used to idealize the fixity of the pier base and/or rotational stiffness of the foundation. If the pier is assumed to be fixed to a rigid foundation, the connection element is assigned a stiffness equal to that of the idealized pier element. The rigid joint elements illustrated in the above figure are used to idealize the geometry and stiffness of the bridge components within the joints. In the analytical model, hinge connection is assumed between the abutment and pile members. An equivalent pile length, le, is assumed to idealize the pile. It is a function of the soil and pile properties and expressed as18,25,26;

where lu is the unsupported length of pile above soil, Ep is modulus of elasticity of pile material, Ip is the moment of inertia of the pile and kh is coefficient of sub-grade reaction of the soil. The pile displacements beyond this equivalent length are negligible. The pile member in the model is therefore assumed to have fixed support conditions at the end. Bridges with large skew angles are not appropriate for integral construction13. Consequently, the above analytical model was not designed to consider skewed bridges. However, integral bridges with skew angles smaller than twenty degrees may be idealized using the proposed model. In the model, frictional forces between the backfill soil and approach slab and wing-walls, resulting from movements due to temperature variations, are also ignored. The above defined model is analyzed for loads numbered from 4 to 14 in the table above. It is noteworthy that the live load applied on the structure must be proportioned to one girder considering the actual transverse distribution of live load effects. The live load transverse distribution factors for slab-on-girder decks can be obtained from bridge design codes21,27. The responses for the deck element obtained from the analysis should be superimposed to those obtained in stage one.
Top of Page

Analyse for the effect of seismic loads A 3-D model is necessary for a realistic representation of the behavior of an integral bridge and load distribution among its various members when it is subjected to seismic loads in

the transverese or longitudinal direction. The following figure shows the analytical model for seismic analysis of a two span integral bridge

The bridge deck is modelled as a 3-D beam element. The in-plane stifness of the deck is relatively much higher than that of other members. Therefore, at the pier location, the bridge deck is modeled in the transverse direction as a rigid bar of length equal to the deck width. This transverse rigid bar is used to model the interaction between the axial deformation of the piers and torsional rotation of the bridge deck, as well as the interaction between the in-plane rotation of the deck and torsional rotation of bridge piers due to unsymmetry or skewness. Elastomeric bearings are also idealized as 3-D beam elements and connected between the cap beam and the transverse rigid bar. If cap beam does not exist, the bearing elements are connected directly to the columns' top. Pin connection is assumed at the joints linking the bearings and the rigid-bar. The product of the elastic modulus, Eib, and moment of inertia, Iib, of the idealized beam element representing the bearings is obtained using the following expression28;

where Gb, Ab and hb are respectively shear modulus, surface area and thickness of the elastomeric bearing. The abutments are idealized as 3-D beam elements. A transverse rigid bar is connected to the end of the abutment. The piles are then idealized as 3-D beam elements and connected to the rigid bar. Pin connection is assumed between the piles and the rigid bar in the

longitudinal direction. The equivalent cantilever model can not be used to idealize the piles since two separate equivalent pile lengths may be obtained for piles with different stiffness in transverse and longitudinal directions. Consequently, the fulll length of the piles is used in the model. The pile elements are divided into a number of equal segments. Then, the lateral stiffness of the soil is calculated at each node level along the pile member using the coefficient of subgrade reaction for the foundation soil. Spring elements with the calculated lateral soil stiffness are then attached to each node. The resistance of the wingwalls to transverse seismic excitations may be taken into account by introducing a spring at the abutment element as shown in Figure 8. The seismically induced soil forces behind the integral bridge abutments may be calculated using the modified Mononobe-Okabe method27,29. The analytical model defined above may be used to conduct a response spectrum analysis or a single mode spectral analysis to obtain the seismic response of the structure using an appropriate site response spectrum.
Top of Page

Design considerations
deck The bridge deck components are designed assuming a continuous frame action at the joints linking the bridge deck to the abutments. A connection detail consistent with the degree of continuity assumed at the joints is provided. The effect of temperature variation and axial compression in the prestressed girders due to backfill soil pressure is considered in the design.
Top of Page

abutment, wingwall and approach slab The abutment is connected monolithically to the deck as shown in Figure 9, to avoid any expansion joint. The abutment height is restricted to the minimum practicle value to reduce the soil pressure and to limit the weight which moves with the deck. However, the minimum penetration required for frost protection is provided. The frost penetration requirement can be reduced to minimize abutment height by providing insulation at the bottom of the abutment. It is recommended that abutments at both sides of the bridge be of equal height since a difference in abutment heights causes unbalanced lateral load resulting in sidesway. Additionally, the soil under the approach slab may be sloped to reduce the height of the soil behind the abutment. This practice is also useful in preventing the compaction of the soil behind the abutment wall due to vehicular traffic. It also reduces the resistance of frictional forces, between the soil and the approach slab, to bridge movement. Turn-back wingwalls parallel to the roadway, carried by the structure, is preferably used. Their size is minimized to allow the substructure to move with minimum resistance. In the province of Ontario, Canada, abutment height and wingwall length is limited to 6 and 7 metres respectively13.

The approach slab is built integral with the abutment to prevent water penetration. An expansion joint is provided at the end of the approach slab. The approach slab is designed as a simply supported structure spanning over the backfill soil behind the abutment to prevent compaction of backfill material.
Top of Page

abutment piles A single row of piles is used to support the abutments. The design of piles may be carried out using the equivalent cantilever method as a beam-column with a fixed base at some distance below the ground surface18,25,26. A pin connection is recommended between the pile top and abutment to allow free rotation of the pile top about an axis perpendicular to bridge longitudinal direction. if the connection is designed as fixed, plastic bending moments may be produced at the pile top due to thermal movements and effect of vehicular traffic3,16-20. In the absence of rigorous theoretical and experimental studies, it may be speculated that the repetitive variation of temperature and the effect of live load may therefore cause low cycle fatigue in steel piles. If the pile supporting system utilizes the frictional forces between the piles and the soil, consideration should be given to the effect of lateral displacement of the piles on the frictional resistance. As the piles will be moving laterally with temperature variations, a gap may be produced between the disturbed soil and the pile. This may result in considerable decrease of the frictional resistance of the piles. Therefore, the piles should be designed using the effective frictional pile length reduced by pile displacements25. If the piles are driven into stiff soils, their longitudinal displacement may somehow be restrained. Pre-drilled oversize holes filled with loose send may be provided to reduce the resistance to lateral movements13,18,20.
Top of Page

bearings, pier and footing The pier is expected to deflect and rock on its foundation when the structure contracts or expands due to temperature variation. Elastomeric bearings of adequate thickness may be used to reduce the flexibility demand of the pier. The bearings are designed to accommodate the movements of the bridge and to support vertical loads coexisting with rotation of the deck. The pier footing is designed as narrow as possible in the longitudinal direction of the bridge to allow partial rotation of the pier at its base. If the footing is supported on piles, the pile group is designed to allow some rotation of the footing. Top of Page

Bibliography

1 Wolde, Tinsae, A. M., Klinger, J. E., White, E.J., 'Performance of Jointless Bridges', ASCE Journal of Performance of Constructed Facilities, Vol. 2, No. 2, 1988, pp 111-128. 2 Wolde, Tinsae, A. M., Klinger, J. E., Mullangi, R., Bridge Deck Joint Rehabilitation or Retrofitting - Final Report, Department of Civil Engineering, Maryland University, College Park, MD, USA, 1988 3 Burke, M. P., Jr., 'Bridge Deck Joints', NCHRP Synthesis of Highway Practice, No 141, Transportation Research Board, National Research Council, Washington, D.C., USA, 1988. 4 Burke, M. P., Jr., 'Integral Bridge Design is on The Rise', AISC Modern Steel Construction, Vol. 30, No. 4, 1990, pp 9-11. 5 Steiger, D. J., 'Jointless Bridges Provide Fuel For Controversy', Roads and Bridges, Vol. 31 (11), 1993, pp 48-54. 6 Burke, M. P., Jr., 'Integral Bridges', Transportation Research Record, No 1275, Transportation Research Board, National Research Council, Washington, D.C., USA, 1990. 7 Soltani, A. A., Kukreti, A. R., 'Performance Evaluation of Integral Bridges', Transportation Research Record, No 1371, Transportation Research Board, National Research Council, Washington, D.C., USA, 1992, pp 17-25. 8 Burke, M. P., Jr., 'Semi-Integral Bridges: Movements and Forces', Transportation Research Record, No 1460, Transportation Research Board, National Research Council, Washington, D.C., USA, 1994, pp 1-7. 9 Burke, M. P., Jr., ' Integral Bridges: Attributes and Limitations', Transportation Research Record, No 1393, Transportation Research Board, National Research Council, Washington, D.C., USA, 1993, pp 1-8. 10 Burke, M. P., Jr., 'The Design of Concrete Integral Bridges', Concrete International, June, 1993, pp 37-42. 11 Hambly, E. C., Nicholson, B. A., 'Prestressed Beam Integral Bridges', The structural Engineer, Vol. 68, No. 23, 1990, pp 474-481. 12 Hayward, A., 'Continuous and Jointless Steel Bridges', Transportation Research Laboratory Record 19, Crowthorne, U.K., 1992, pp 83-90. 13 Husain, I., Bagnariol, D., Integral Bridges, Ontario Ministry of Transportation, Report SO-96-01, St. Catharines, Ontario, Canada, 1996. 14 Hamley, E. C., 'Integral Bridge Abutment Details in Practice and Theory', Transport Research Laboratory Record 19, Crowthorne, U.K., 1992. 15 Federal Highway Administration (FHWA), Seismic Design and Retrofit Manual for Highway Bridges, FHWA-IP-87-6, U.S. Department of Transportation, Federal Highway Administration, Washington, D. C., USA, 1987. 16 Emanual, J. H., Hulsey, J. L., Best, J. L., Senne, J. H., Thompson, L. E., Current Design Practice for Bridge Superstructures Connected to Flexible Substructures, Civil Engineering Study 73-3, University of Missouri-Rolla, Missouri, 1973.

17 Loveall, C. L., 'Jointless Bridge Decks', Civil Engineering, November, 1985. 18 Abendroth R. E., Greimann, L. F., 'A Rational Design Approach for Integral Bridge Piles', Transportation Research Record, No 1223, National Research Council, Washington, D.C., USA, 1989, pp 12-23. 19 Abendroth R. E., Greimann, L. F., Ebner, P. B., 'Abutment Pile Design for Jointless Bridges', ASCE J. Struct. Engrg., 115 (11), 1989, pp 2914-2929. 20 Girton, D. D., Hawkinson, T. R., Greimann, L. F., 'Validation of Design Recommendations for Integral-Abutment Piles', ASCE Journal of Structural Engineering, Vol. 117, No. 7, 1991, pp 2117-2134. 21 Ontario Highway Bridge Design Code, Third Edition, Ministry of Transportation, Quality and Standards Division, Ontario, Canada, 1991. 22 Barker, R. M., Duncan, J. M. K., Rojiani, K. B., Ooi, P. S. K, Kim, S.G., Manuals For The Design of Bridge Foundations, NCHRP Report 343, Transportation Research Board, National Research Council, Washington, D.C., USA, 1991. 23 Clough, G. M., Duncan, J. M., Foundation Engineering Handbook, Second Edition, Edited by H. Y., Fang, Van Nostrand Reinhold, New York, NY, 1991. 24 Roeder, C., W., Moorty, S., 'Thermal Movements in Bridges', Transportation Research Record, No 1290, Transportation Research Board, National Research Council, Washington, D.C., USA, 1990, pp 135-143. 25 Greimann, L. F., Abendtroth, R. E, Johnson, D. E., Ebner, P. B., Pile Design and Tests for Integral Bridges, Final Report, Iowa Department of Transportation, Project HR-273, 1987. 26 Girton, D. D., Hawkinson, T. R., Greimann, L. F., Bergenson, K., Ndon, U., Abendorth, R. E., Validation of Design Recommendations for Integral Piles, Iowa Department of Transportation, Project HR-292, 1989. 27 AASHTO LRFD Bridge Design Specifications, First Edition, American Association of State Highway Transportation Officials, Washington, D. C., USA, 1994. 28 Dicleli M., Effects of Extreme Gravity and Seismic Loads on Short to Medium Span Slab-On-Girder, Steel Highway Bridges, PhD Thesis, Department of Civil Engineering, University of Ottawa, Ontario, Canada, 1993. 29 Demetrios, E. T., Bridge Engineering; Design, Rehabilitation and Maintenance of Modern Highway Bridges, McGraw-Hill, New York, NY, USA, 1995. 30 Handbook of Steel Construction, Fifth Edition, Canadian Institute of Steel Construction, Willowdale, Ontario, Canada, 1993. 31 Ontario Modular Bridge Analysis System, Release 6.3, Transportation Engineering Divison, Structural Office, St. Catharines, Ontario, Canada, 1998.

Top of Page

web hosting domain names web design online games

You might also like