Sie sind auf Seite 1von 9

ACI STRUCTURAL JOURNAL

Title no. 99-S60

TECHNICAL PAPER

Computer-Based Tools for Design by Strut-and-Tie Method: Advances and Challenges


by Tjen N. Tjhin and Daniel A. Kuchma
The strut-and-tie method (STM) is gaining recognition as a codeworthy and consistent methodology for the design of D- (discontinuity) regions in structural concrete. Unfortunately, the development of code provisions for the STM has been hampered by uncertainties in defining the strength and dimensions of the idealized load-resisting truss (or strut-and-tie model). In addition, the has been encumbered by an iterative and time-consuming design procedure in which many geometric details need to be considered. To overcome this problem, researchers are developing computer-based design tools, including the authors computer-aided strut-and-tie (CAST) design tool. CAST provides a graphical working environment for all aspects of the design process, including definition of the D-region, selection of the strut-and-tie model, truss analysis, member definitions, and creation of a design summary. This study reports on the STM, the barriers to its advancement, the capabilities of computer-based design tools, and the CAST program. It also makes suggestions for future STM research.
Keywords: structural concrete; strut; tie.

Beginning with a brief description of the STM, this paper discusses complications in the STM design process and the role that computer-based tools can serve in overcoming these obstacles. Then, a summary of the capabilities of a few computer-based STM design tools is presented, with emphasis on the features that distinguish one tool from another. This includes work at Purdue University, the Swiss Federal Institute of Technology (ETH), the University of Stuttgart, and others. Following this, the computer-aided strut-and-tie (CAST) design tool that is being developed by the authors is presented. The study concludes with a summary of the challenges that lie ahead for the STM and associated computer-based design tools. RESEARCH SIGNIFICANCE Because of the inadequacy of traditional code provisions and detailing practices, most structural problems occur in D-regions. The STM has the potential to provide a consistent, well-founded, and widely applicable design methodology for D-regions, but it is marred by a cumbersome hand-based design process. To overcome this problem, computer-based design and analysis tools that bring simplicity and transparency to the STM design process, and thus can improve how D-regions are designed, are being developed. This paper presents the role and capabilities of these programs, summarizes uncertainties in the STM design methodology, and suggests directions for future research. STM FOR DESIGN OF D-REGIONS Background The idea of the STM came from the truss analogy method introduced independently by Ritter and Mrsch approximately 100 years ago for the shear design of B-regions. The truss analogy, or truss model, was used to idealize the flow of forces in a cracked concrete beam. In parallel with the increasing availability of experimental results and the development of limit analysis in the plasticity theory, the truss analogy method has been validated and improved considerably in the form of full member or sectional design procedures. The truss model has also been used as a basis for torsion-design methods. An excellent summary of the development of truss model for shear design of B-regions can be found in Reference 5. The STM was developed after Schlaich, Schfer, and Jennewein3 extended the use of a truss model to D-regions.
ACI Structural Journal, V. 99, No. 5, September-October 2002. MS No. 01-236 received August 6, 2001, and reviewed under Institute publication policies. Copyright 2002, American Concrete Institute. All rights reserved, including the making of copies unless permission is obtained from the copyright proprietors. Pertinent discussion will be published in the July-August 2003 ACI Structural Journal if received by March 1, 2003.

INTRODUCTION In selecting the appropriate design approach for structural concrete, it is useful to classify portions of the structure as either B- (beam or Bernoulli) regions or D- (disturbed or discontinuity) regions. B-regions are those parts of a structure in which it is reasonable to assume that there is a linear variation in strain over the depth of the section. D-regions are the remaining parts of the structure in which there is a complex variation in strain, occurring near abrupt changes in geometry (geometrical discontinuities) or concentrated forces (statical discontinuities). Based on St. Venants principle, the extent of a D-region spans approximately one section depth of the region on either side of the discontinuity. The distinction between B- and D-regions is illustrated in Fig. 1. Most design practices for B-regions are based on a model for behavior. For example, the design for flexure is based on conventional beam theory while the design for shear is based on the well-known parallel chord truss analogy. In contrast, the most familiar types of D-regionssuch as deep beams, corbels, joints, and pile capsare principally designed by empirical approaches, such as those given in ACI 318-99,1 or by using common detailing practices. For most other types of D-regions, code provisions provide little guidance to designers. Not surprisingly, most structural problems occur in D-regions. The strut-and-tie method2-4 (STM) is emerging as a codeworthy methodology for the design of all types of D-regions in structural concrete. Unfortunately, this conceptually powerful method can be complicated by the need to perform time-consuming calculations and graphical procedures. It is for this reason that computer-based graphical design aids are being developed. 586

ACI Structural Journal/September-October 2002

ACI member Tjen N. Tjhin is a doctoral candidate in the Department of Civil and Environmental Engineering at the University of Illinois at Urbana-Champaign. His research interests include nonlinear analysis and design of concrete structures. ACI member Daniel A. Kuchma is an assistant professor of civil and environmental engineering at the University of Illinois at Urbana-Champaign. He is a member of ACI Subcommittee 318-E, Shear and Torsion; and Joint ACI-ASCE Committee 445, Shear and Torsion.

Nodes are analogous to joints in a truss, and are where forces are transferred between struts and ties. As a result, these regions are subject to a multidirectional state of stress. Nodes are classified by the types of forces being connected. Figure 4 shows basic types of nodes; in the figure, C is used to denote compression and T is used to denote tension. STM design process The STM design process involves several steps that are illustrated in Fig. 5 using the design example of a dappedended beam, and are described as follows: 1. Defining the boundaries of the D-region and then evaluating the concentrated, distributed, and sectional forces that act on the boundaries of this region; 2. Sketching a strut-and-tie model and solving for the truss member forces; 3. Selecting the reinforcing or prestressing steel that is necessary to provide the required tie capacity and ensuring that this reinforcement is properly anchored in the nodal zones (joints of the truss); 4. Evaluating the dimensions of the struts and nodes such that the capacity of these components is sufficient to carry the design force values; and 5. Providing distributed reinforcement to increase the ductility of the D-region. Because equilibrium of the truss with the boundary forces must be satisfied (Step 2) and stresses everywhere must be below defined code limits (Steps 3 and 4), the STM is a lowerbound (static or equilibrium) method of limit analysis.

Strut-and-tie models In the STM, the complex flow of internal forces in the D-region under consideration is idealized as a truss carrying the imposed loading through the region to its supports. This truss is called a strut-and-tie model. Like a real truss, a strutand-tie model consists of struts and ties interconnected at nodes (nodal zones or nodal regions). A selection of strut-and-tie models for a few typical D-regions is illustrated in Fig. 2. Struts are the compression members of a strut-and-tie model and represent concrete stress fields whose principal compressive stresses are predominantly along the centerline of the strut. As shown in Fig. 2, struts are usually symbolized using a broken line. The actual shape of a strut, however, can be prismatic, bottle-shaped, or fan-shaped (Fig. 3). Struts can be strengthened by steel reinforcement and, when this is the case, they are called reinforced struts. Ties are the tension members of a strut-and-tie model. Ties mostly represent reinforcing steel, but they can occasionally represent prestressing steel or concrete stress fields with principal tension predominantly in the tie direction. Ties are usually denoted using a solid line.

Fig. 1Example of division of B- and D-regions in common structure.

Fig. 2Examples of strut-and-tie models. ACI Structural Journal/September-October 2002 587

STM design provisions STM design provisions consist of rules for defining the dimensions and ultimate stress limits of struts and nodes as well as the requirements for the distribution and anchorage
fc1<fc

fc

fc

of reinforcement. Guidelines6,7 for STM design have been developed for European practice. Provisions for the STM were incorporated in the Canadian Concrete Design Code8 in 1984 and in the AASHTO LRFD Bridge Design Specifications9 in 1994. Another specific set of provisions has been developed to be included as an alternative design procedure in Appendix A, Strut and Tie Models, of ACI 318-02.10 Due to uncertainties associated with defining the characteristics of an idealized truss within a continuum of structural concrete, there are substantial differences in the rules used in these provisions and guidelines. COMPLICATIONS IN DESIGN USING STM Although the STM is conceptually simple, numerous complications can encumber the five-step design process, and they are the motivation for the development of computerbased tools. A few of these complications are briefly described as follows. Selection of appropriate strut-and-tie models Because the STM is a lower-bound design methodology, more than one admissible truss can be developed for each load case as long as the selected truss is in equilibrium with the boundary forces and the stresses in the struts, ties, and nodes of the truss are within acceptable limits. As a result of limited ductility in the structural concrete, however, there are only a small number of viable solutions for each design region. Figure 6 illustrates the ways in which some solutions are preferable to others.

fc (a)

fc (b)

fc (c)

Fig. 3Basic type of struts: (a) prismatic; (b) bottleshaped; and (c) fan-shaped (adapted from Schlaich, Schfer, and Jennewein).3

Fig. 4Basic type of nodes: (a) CCC; (b) CCT; (c) CTT; and (d) TTT.

Fig. 5Steps in STM design process. 588 ACI Structural Journal/September-October 2002

The most widely used guidance for selecting the appropriate truss so far is probably that suggested by Schlaich, Schfer, and Jennewein,3 who proposed arranging the truss members within 15 degrees of the principal stress trajectories obtained from a linear elastic solution. With this approach, both serviceability and ultimate limit states are expected to be satisfied. Based on the minimum strain-energy principle, Schlaich, Schfer, and Jennewein also suggest selecting a truss in which the total length of ties is a minimum. This guidance is very helpful, but elastic solutions or solutions for minimum length of ties are not always easy to obtain. The other method includes recent work by Ali and White.11 Also using the energy approach, they introduced the elastic strain compatibility error (SCER) concept to measure how well a selected strut-and-tie model deviates from the elastic solution; the lower the SCER value, the closer it is to the elastic stress distribution. Adjustment of truss geometry and dimensions The initially selected geometry of the truss, including the strut and node dimensions, must often be adjusted to satisfy stress limit criteria, to fit the struts and nodes within the D-region boundaries, to investigate other configurations, and to optimize the design. Refinement of the truss geometry sometimes needs to be undertaken for similar reasons. This can make hand-based solutions prohibitively time-consuming, particularly for the design of complex D-regions. Geometry and dimensions of ties and nodes To determine the dimensions and shape of a node, it is necessary to know the widths of the incoming struts and ties. The width of a strut is typically selected so that a particular code-specified stress limit value is not exceeded. It is less clear, however, how to define the effective width of ties as well as the dimensions and shape of nodal zones. The classic method of dimensioning a node is by arranging the node shape so that the stresses on all sides of the node from the truss member forces as well as from the boundary forces meeting at the nodeare equal. The biaxial state of stress inside the node is hydrostatic; in other words, the in-plane stresses are isotropic, homogeneous, and equal to those on the sides. Arranging the node in this shape can be done by sizing the boundaries of the node so that they are proportional and perpendicular to the forces acting on them (Fig. 7). In defining the width of a tie, the tie force is treated as a compressive force acting from behind the node (Fig. 7(b) through (d)). This type of node is called a hydrostatic node. Laying out nodes in this manner can be very laborious, especially for nodes bounded by more than three truss members (Fig. 8(a)), as the centerlines of truss members framing into hydrostatic nodes of more than three sides are unlikely to coincide (Fig. 8(b)). A more simplified method that is applicable for nodes with typical configurations was proposed by Schlaich and Schfer.4 The shape of a node is simply formed by the intersection of actual dimensions of struts and ties whose centerlines coincide at that node (Fig. 8(c)). The in-plane stresses acting on all sides of the node do not need to be equal, but the stress on each side must be constant and below the node stress limit. This convenient method, however, must be used discriminately for complex node configurations. Statically indeterminate strut-and-tie models There is little guidance available for evaluating the relative stiffness of members in a statically indeterminate strut-andACI Structural Journal/September-October 2002

Fig. 6Two admissible strut-and-tie models for squat structural wall under horizontal force: (a) workable truss; and (b) less feasible truss due to excessive ductility demands.
fc fc fc fc fc (a) fc T fc fc fc T (b) T fc fc fc

fc fc fc

fc fc fc (c) T T fc (d) T

Fig. 7Examples of hydrostatic nodes: (a) CCC; (b) CCT: (c) CTT; and (d) TTT. tie model. As a result, the designer is unsure how to determine the distribution of forces in these types of trusses. The classic way to handle a statically indeterminate case is to employ the so-called plastic truss method. In this method, the most heavily loaded ties in the truss are assumed to have yielded so that the tire forces become known and the truss becomes statically determinate. This method, however, must be used with caution because of strain compatibility requirements and limited ductility in concrete. Another method, suggested by Schlaich and Schfer,4 is to decompose the statically indeterminate truss into several statically determinate trusses. Again, as a result of the strain compatibility requirements, this technique requires a reasonable estimation of the stiffness and the imposed-loading distribution of each statically determinate truss. Multiple load cases and load combinations The need to consider multiple load cases and load combinations can increase the time required to complete a design using the STM by several times. This is so because different strut-and-tie models may need to be prepared to handle 589

C1 fc fc fc fc

C2

C3 fc 2

fc fc

fc

C4 (a)

fc
fc1 fc3

fc fc fc

fc (b)

fc4 (c)

fc
Fig. 9Arrangement of CCC node of Fig. 8(a) that comprises two hydrostatic nodes of triangular shape and short prismatic strut. capabilities have also been created.19,20 The latter analysis may be particularly useful when the capacity of a D-region using a strut-and-tie model is to be assessed. The use of principal stress trajectories for guiding the construction of strut-and-tie models has also been extended to the automatic generation of strut-and-tie models. An example of this can be seen in the work by Harisis and Fardis,22 who used statistical analysis of principal stress data obtained from linear finite element analysis to identify locations of struts and ties. The strut-and-tie models generated using this approach consist of triangles. Another example is the work by Rckert.14 Based on the fact that regions bounded by the mesh of principal stress trajectories represent finite elements subjected only to normal stress on each side, an strut-and-tie model can be formed using these finite elements. Ali and White11 recently developed algorithms for the automatic generation of strut-and-tie models. Based on user-selected locations of nodes, an optimal truss solution can be generated following the elastic, minimum reinforcement volume, or composite criterion. Approaches in nodal-zone construction One simple computer-based approach to construct a nodal zone is to define the node shape as the intersection of the dimensions of truss members whose centerlines coincide, as illustrated in Fig. 8(c). The adequacy of the node is checked using linear elastic finite element analysis with the Coulomb failure criterion18 or nonlinear finite element analysis with failure criteria determined from experimental test data of two-dimensional plain concrete.19,20 Another approach is to use modified hydrostatic node construction.12 In this approach, a node with more than three members intersecting is handled by breaking down the node into several hydrostatic nodes of triangular shapes connected by short prismatic struts. Figure 9 shows how the CCC node of Fig. 8(a) is arranged using this approach. Trapezoidal transition stress zones between a node and the intersecting truss members were also formulated using this approach, allowing different stress intensities of truss members to act on the node. The length of these transition zones is determined using separation failure criteria of modified Coulomb material. ACI Structural Journal/September-October 2002

Fig. 8Example of CCC node with four truss members intersecting: (a) forces acting on node; (b) node shape arrangement causing hydrostatic stress state; and (c) simplified arrangement. each different loading situation. In addition, load cases usually cannot be superimposed directly to form load combinations as a result of strain compatibility requirements. To summarize the aforementioned complications, the STM design process can be encumbered by the challenge of selecting the initial truss, the need to iteratively adjust and refine truss geometry and member dimensions, the lack of guidance for the design of statically indeterminate trusses, and the need to consider multiple load cases and load combinations. Additionally, the STM does not yet address structural performance under service loading. Computer-based strutand-tie design programs, as described in the following section, can overcome many of these challenges. ADVANCES IN COMPUTER-BASED STM Most of the work on computer-based STM in Europe was done at the Swiss Federal Institute of Technology (ETH) under the supervision of Anderheggen12,13 and at the University of Stuttgart under the supervision of Schlaich and Schfer.14,15 In North America, a series of computer-based STM tools were developed by Ramirez and his students at Purdue University 16-20 as well as by Ali and White11 at Cornell University. In general, all of the tools that have been developed in those institutions offer interactive analyses of STM. The construction and modification of the boundaries of the structure under consideration as well as the selected strut-and-tie models are done graphically using a mouse and keystrokes. The main thing that distinguishes one tool from another is the approach they use to obtain valid STM solutions, as well as other features that they provide. Some of the noteworthy capabilities provided by these STM programs are presented as follows. Tools for strut-and-tie model construction As discussed previously, principal stress trajectories obtained from a linear elastic solution can provide guidance in selecting an appropriate strut-and-tie model. For this purpose, many computer-based STM tools are equipped with a linear elastic finite element analysis that can be used for generating plots of principal stress trajectories.17,18,21 These plots are then used as a backdrop for constructing a strut-and-tie model. Programs that include nonlinear finite-element analysis 590

Fig. 10User interface of CAST education and design tool. In determining the width of a tie, both of these approaches treat the tie as a compressive force that acts from behind the node. Procedures for determining truss member forces Several computer-based procedures have been developed to determine truss member forces, especially for statically indeterminate cases. The simplest procedure is to use an elastic truss analysis. This analysis is not very convenient, because the relative stiffness of truss members, which is very difficult to determine, must be supplied. Anderheggen and Schlaich13 developed two procedures based on a lower-bound solution: rigid-plastic optimal design and shakedown optimal design. In the first procedure, the member forces of a statically indeterminate strut-and-tie model are determined in such a way that minimum tie resistances, corresponding to the minimum weight of steel ties, are obtained. This is done through linear programming. Determining the distribution of forces in this manner is possible because strain compatibility equations are ignored in the lowerbound solution. Based on the static shakedown theorem, the second procedure is similar to the first, except that the elastic truss solutions have to be considered. Because the computational time is reduced, this procedure is useful when several load cases need to be considered. The obtained solution lies between the elastic solution and the rigid-plastic solution. Prediction of load-deformation response and capacity Work on computer-based STMs has been advanced to predict load-deformation response and capacity using strut-and-tie models.14,15 In this work, the load-deformation response of struts and ties must be specified, and the minimum energy criterion is employed. For each loading stage, the geometry of the strut-and-tie model is adjusted so that the internal energy of the system is at a minimum. The predictions of these methods are reported by the authors of these programs to show good agreement with experimental data. ACI Structural Journal/September-October 2002 CAST DESIGN TOOL Overview The development of the CAST program was motivated by the desire to bring efficiency and transparency to the STM design process. To this end, the goal was to create a computer-based graphical working environment in which the designer is able to readily sketch the boundaries of the D-region, draw a proposed internal load-resisting truss, seamlessly solve for member forces, readily select the dimensions of struts and reinforcement for ties, and then generate a printout that summarizes the design. Other necessary features included the ability to make on-screen adjustments of all key variables, tailor the program for specific design code provisions, and handle multiple load cases and combinations. In developing these capabilities, it was considered essential that CAST functioned as a transparent tool that helped the designer to explore and evaluate potential solutions. CAST is believed to be well on its way to achieving these objectives and thereby enabling the STM to achieve its potential of providing engineers with an efficient, consistent, and well-founded methodology for the design of D-regions in structural concrete. Basic operation of CAST A user begins by defining the boundaries of the D-region using a mouse, by numerical entry, or by selecting an object from a group of templates. The position of all of the nodes that define the shape of the object can be easily adjusted at any stage in the design process. The applied loading, bearing plates, material properties, and support conditions can then be selected from toolbars or specified in dialogue boxes. Next, the proposed load-resisting truss can be drawn as a series of straight lines using the mouse. The truss can then be analyzed and the results displayed alongside the truss members. Using a toolbar, the user can then select and position the estimated tie reinforcement as well as select the widths of the struts. The capacity of the struts and ties as well as the applied stress on the faces of the nodes can then be displayed. The designer can use the nodal zoom-edit tools to change features of the node that influence node and truss geometry, and thereby 591

(f) (d) (e) (c) (a) (b)

Fig. 11Types of struts in discontinuity region: (a) prismatic in uncracked field; (b) prismatic in cracked field where struts are parallel to cracks; (c) prismatic in cracked field where struts are not parallel to cracks; (d) bottle-shaped with crack control reinforcement; (e) bottle-shaped without crack control reinforcement; and (f) confined strut. affect the stresses on the face of the nodes and design strengths. Once the user is satisfied with the design, the detailed designincluding member forces, capacities, stresses on the faces of nodal zones, and reinforcement detailscan be printed. Figure 10 provides an image of CASTs working environment. The interior images show a view of the entire structure with applied loads and support conditions, a detail of the labeled forces, the selection of tie reinforcement, and the detail of a node. REMAINING CHALLENGES TO BE RESOLVED IN STM Despite the advances in computer-based STM tools, a number of uncertainties still remain in the STM. These uncertainties must be addressed in future research so that CAST or other computer-based tools can make the strut-and-tie design process as efficient and transparent as the STM is conceptually powerful. A few of the main challenges are described as follows. Capacity of struts There is still much debate over the effective compressive strength of a strut. This is reflected in the different strength values specified in codes and guidelines6-10 and proposed by researchers.2-4,23 Nevertheless, it is generally agreed that the strut strength is a fraction of the uniaxial concrete compressive strength as obtained from cylinder tests. The following factors have been identified as influences to the ultimate compressive stress capacity of struts: 1. Shape of strutIf the stress trajectories in a strut are parallel as shown in Fig. 3(a) and 11(a) and (f), the strength of the strut is close to that of the compressive strength of a concrete cylinder. If the strut is within the core of the D-region, however, the compressive stresses in the strut tend to spread out as they move away from the nodes (Fig. 3(b) and 11(d) and (e)). This bottle-shaped stress path can lead to the splitting of a strut at a compressive stress that is considerably less than the cylinder compressive strength; 2. Disturbances in strutOther factors that influence the splitting strength of a strut include initial cracks parallel or inclined to the strut axis and tensile transverse stress or strain induced by a crossing tie or another effect (Fig. 11(b) and (c)); 3. Distributed reinforcementThe use of distributed reinforcement can control splitting of the strut due to spreading or disturbances (Fig. 11(d)). It may also increase the overall structural ductility and thus help improve the performance of a D-region designed using a less-than-ideal truss selection; 4. ConfinementThe performance of the strut can also be enhanced by confinement provided by either a specially 592 designed confining reinforcement or by mass reinforced concrete that surrounds the strut (Fig. 11(f)); and 5. Angle of strut It is well documented24 that the compressive strength of concrete decreases as the level of transverse tensile strain increases. The implication for strutand-tie modeling is that, as the angle between the strut and a tie in a nodal region decreases, the compressive strength of the strut is expected to decrease. Thus, in the design of a deep beam, a steeper strut is expected to be stronger than a horizontal strut. While many of the factors that can influence the compressive strength of concrete struts have been identified, the effect of these factors has yet to be adequately quantified. This is due to the very limited amount of experimental research25 that has been conducted to understand how to define the shape, stiffness, and strength characteristics of struts. Load-deformation response of struts and ties Another major challenge is to accurately estimate the loaddeformation response of struts and ties. This is important for the following reasons: 1. If the stiffness characteristics of the struts and ties are known, the load distribution in statically indeterminate strutand-tie models, as mentioned previously, may be predicted. This is illustrated in Fig. 12, where the point load is considered to be transferred to the support by two distinct load paths: (i) a direct strut from the point of loading to the support; and (ii) a path consisting of two steeper struts connected by a steel tie. The portion of the load taken by each load path will be in proportion to the relative stiffness of these paths; 2. If the stiffness characteristics of the struts and ties at service loads are known, the equivalent deflection response may be predicted for the evaluation of serviceability limit state; and 3. As illustrated in Fig. 12, a tie may cross the path of a strut. The straining and cracking induced in this situation may influence the strength of the strut. By examining the factors that influence the tension-stiffening effect and distribution of cracking in ties, the capacity and response of struts can be better understood. Anchorage and distribution of tie reinforcement In the STM, the selected steel ties can fully develop tensile forces and transfer the forces in the nodes only if the ties are properly anchored in the nodes. There are still uncertainties, however, about the anchorage requirements and the need to distribute the tie reinforcement throughout the nodal region. ACI Structural Journal/September-October 2002

(i) (ii) (i) (ii)

Fig. 12Statically indeterminate strut-and-tie model in simply supported deep beam. The example provided in Fig. 13 illustrates that the distribution and anchorage of tie reinforcement clearly influence the ability to transfer the horizontal component of a diagonal strut to the tie at the end of a simply supported member. The previous research in this area26,27 is very limited. Some known and suggested parameters for future research may include end anchorage of bars (such as straight, hooked, or headed bar, and end plate), bar size and roughness, lateral and vertical spacing between bars, angle of incoming strut, width of bearing plate in nodal zone, use of confinement, length of bar, and use of fibers. Size, shape, and strength of complex nodal zones The anchorage detail examined in the preceding section illustrated some of the complexities of load transfer in nodal regions. As mentioned previously, these regions may have a large variation in their configurations and thus become quite difficult to understand. Previous research in this area28,29 is also very limited, and more research is essential. Suggested factors to study the dimensioning of nodal zones may include the type of truss members (struts or ties), number of intersecting truss members, distribution of tie reinforcement, confinement and use of fibers, level of transverse straining, volume and condition of surrounding concrete, and anchorage conditions of ties. To obtain simplicity in the STM design process, these studies should focus on determining the shape of nodal zones in which the centerlines of incoming struts and ties coincide. Serviceability limit state It is recognized that the design of structural concrete should consider both ultimate and serviceability limit states. Until now, the development of STM has been focused on the ultimate limit state; the serviceability limit state is only implicitly considered through the selection of appropriate strut-and-tie models. Evaluation of service load deflections by the STM requires the selection of strut-and-tie model configuration and the estimation of axial truss member stiffness characteristics that give equivalent response to the deflections of the real structure. The evaluation of crack widths at service loads requires reasonable estimates of effective concrete area in tension around the ties of reinforcing steel. ACI Structural Journal/September-October 2002

Fig. 13Examples of various tie anchorage conditions. SUMMARY The STM offers a rational and consistent basis for the design of D-regions in structural concrete. This is a significant achievement, as typical design practice for D-regions in structural concrete consists of inadequate empirical design provisions and detailing practices. The main steps in the STM design process involve defining the D-region and the boundary forces acting on the region, visualizing a truss carrying the boundary forces in the D-region (that is, the strut-and-tie model), solving for the truss member forces, providing reinforcement to serve as the steel ties, dimensioning the struts and nodes, and providing distributed reinforcement for ductility. Unfortunately, there are a number of complexities and challenges to designing with the STM that are limiting its use in practice. These include the need to select a feasible idealized truss, the need to iteratively adjust and refine the truss geometry and dimensions, the need to consider multiple load cases, uncertainties in the dimensioning of truss components, and the lack of guidance for the design of statically indeterminate trusses. An overriding concern is the time required to complete an STM design by hand and the opportunities for error that are introduced by the use of a time-consuming hand-based design process. Advances in computer-based tools, such as CAST, have been made to overcome the complications and challenges in the STM design process. Different approaches and specific features have been implemented to obtain valid STM solutions. These include tools for generating strut-and-tie models, approaches to constructing nodal regions, procedures for determining truss member forces, and prediction of loaddeformation response and capacity. Uncertainties in defining the dimensions, stiffness, and strength of struts, ties, and nodes are major barriers to the development of improved STM design provisions and computer-based design aids. Significant experimental research is required to better understand the performance of individual struts, ties, and nodes, as well as both simple and complex strut-and-tie models. In future research, it is essential that very detailed and accurate measurements be made of surface deformations to properly assess the structural behavior. Some of this required research is being conducted as part of the NSF award received by the second author, but significantly more research is required to address the many areas of uncertainty and to help the STM become a reliable and 593

cost-effective methodology for the design of all types of D-regions. Additional information about the STM and the CAST design tool, including detailed design examples, can be found on the authors strut-and-tie resources webpage at www.ce.uiuc.edu/kuchma/strut&tie. A recent paper by the authors30 compares the STM provisions used in the AASHTO-LRFD and ACI 318-02 specifications and provides a design example using CAST. ACKNOWLEDGMENTS
The CAST program has been under development since 1998, and it will run on WindowsTM operating platforms. The current version of this program is freely available from http://www.cee.uiuc.edu/kuchma/strut&tie. The development of CAST has been supported by the University of Illinois and the Portland Cement Association. The continued development of CAST will be supported by an award from the National Sciences Foundation through March of 2006. This provides an exceptional opportunity to create a freely available and sophisticated design tool that will be developed using substantial input from the structural engineering community. Practicing engineers, students, and educators are encouraged to visit the strut-and-tie website, to download the CAST program, and to provide feedback and suggestions for its continued development. From this website, detailed information about the STM, completed design examples, tutorials, instructional modules, and other design and educational resources are being provided.

NOTATION
C, C1, C2, ... = fc, fc1, fc2, ... = h, h1, h2, ... = T = compression force in strut-and-tie model normal stress in concrete depth of structural member tension force in strut-and-tie model

REFERENCES
1. ACI Committee 318, Building Code Requirements for Structural Concrete (ACI 318-99) and Commentary (318R-99), American Concrete Institute, Farmington Hills, Mich., 1999, 391 pp. 2. Marti, P., Basic Tools of Reinforced Concrete Beam Design, ACI JOURNAL, Proceedings V. 82, No. 1, Jan.-Feb. 1985, pp. 46-56. 3. Schlaich, J.; Schfer, K.; and Jennewein, M., Toward a Consistent Design of Structural Concrete, Journal of the Prestressed Concrete Institute, V. 32, No. 3, May-June 1987, pp. 74-150. 4. Schlaich, J., and Schfer, K., Design and Detailing of Structural Concrete Using Strut-and-Tie Models, The Structural Engineer, V. 69, No. 6, Mar. 1991, pp. 113-125. 5. Joint ACI-ASCE Committee 445, Recent Approaches to Shear Design of Structural Concrete, Journal of Structural Engineering, ASCE, V. 124, No. 12, Dec. 1998, pp. 1375-1417. 6. CEB-FIP Model Code 1990, Thomas Telford Services, Ltd., London, for Comit Euro-International du Bton, Lausanne, Switzerland, 1993, 437 pp. 7. FIP Commission 3, FIP Recommendation 1996, Practical Design of Structural Concrete, Fdration Internationale de la Prcontrainte, May 1998, 142 pp. 8. CSA Technical Committee on Reinforced Concrete Design, A23.3-94: Design of Concrete Structures, Canadian Standards Association, Rexdale, Ontario, Dec. 1994, 199 pp. 9. American Association of State Highway and Transportation Officials, AASHTO-LRFD Bridge Specification, First Edition, Washington, D.C., 1994, 1091 pp. 10. ACI Committee 318, Building Code Requirements for Structural Concrete (ACI 318-02) and Commentary (318R-02), American Concrete Institute, Farmington Hills, Mich., 2002, 443 pp. 11. Ali, M. A., and White, R. N., Automatic Generation of Truss Model

for Optimal Design of Reinforced Concrete Structures, ACI Structural Journal, V. 98, No. 4, July-Aug. 2001, pp. 431-442. 12. Schlaich, M., and Anagnostou, G., Stress Fields for Nodes of Strutand-Tie Model, Journal of Structural Engineering, ASCE, V. 116, No. 1, Jan. 1990, pp. 13-23. 13. Anderheggen, E., and Schlaich, M., Computer Aided Design of Reinforced Concrete Structures using the Truss Model Approach, Proceedings of the Second International Conference on Computer Aided Analysis and Design of Concrete Structures, N. Bicanic and H. Mang, eds., Zell am See, Austria, Apr. 1990, pp. 539-550. 14. Rckert, K. J., Design and Analysis with Strut-and-Tie Models Computer-Aided Methods, Structural Concrete, IABSE Colloquium, Stuttgart, International Association for Bridge and Structural Engineering, Zrich, Mar. 1991, pp. 379-384. 15. Sundermann, W., and Mutscher, P., Nonlinear Behaviour of Deep Beams, Structural Concrete, IABSE Colloquium, Stuttgart 1991, International Association for Bridge and Structural Engineering, Zrich, Mar. 1991, pp. 385-390. 16. Benabdallah, S.; Ramirez, J. A.; and Lee, R. H., Computer Graphics in Truss-Model Design Approach, Journal of Computing in Civil Engineering, ASCE, V. 3, No. 3, July 1989, pp. 285-301. 17. Alshegeir, A., and Ramirez, J. A., Analysis of Disturbed Regions with Strut-and-Tie Models, Publication No. CE-STR-90-1, School of Civil Engineering, Purdue University, West Lafayette, Ind., 1990, 85 pp. 18. Alshegeir, A., and Ramirez, J. A., Computer Graphics in Detailing Strut-Tie Models, Journal of Computing in Civil Engineering, V. 6, No. 2, Apr. 1992, pp. 220-232. 19. Yun, Y. M., and Ramirez, J. A., Strength of Struts and Nodes in Strut-Tie Model, Journal of Structural Engineering, ASCE, V. 122, No. 1, Jan. 1996, pp. 20-29. 20. Yun, Y. M., Computer Graphics for Nonlinear Strut-Tie Model Approach, Journal of Computing in Civil Engineering, ASCE, V. 14, No. 2, Apr. 2000, pp. 127-133. 21. Mish, K.; Nobari, F.; and Liu, D., An Interactive Graphical Strutand-Tie Application, Proceedings of the Second Congress on Computing in Civil Engineering, J. P. Mohsen, ed., American Society of Civil Engineers, New York, 1995, pp. 788-795. 22. Harisis, A., and Fardis, M. N., Computer-Aided Automatic Construction of Strut-and-Tie Models, Structural Concrete, IABSE Colloquium, Stuttgart, International Association for Bridge and Structural Engineering, Zrich, Mar. 1991, pp. 533-538. 23. MacGregor, J. G., Reinforced Concrete: Mechanics and Design, Prentice Hall, Inc., Englewood Cliffs, N. J., Third Edition, 1997, 939 pp. 24. Vecchio, F. J., and Collins, M. P., The Modified Compression-Field Theory for Reinforced Concrete Elements Subjected to Shear, ACI JOURNAL, Proceedings V. 83, No. 2, Mar.-Apr. 1986, pp. 219-231. 25. Joint ACI-ASCE Committee 445, Strut-and-Tie Bibliography (Master List), American Concrete Institute, Farmington Hills, Mich., Sept. 1997, 50 pp. 26. Lee, D. D. K., An Experimental Investigation in the Effects of Detailing on the Shear Behaviour of Deep Beams, MASc thesis, Department of Civil Engineering, University of Toronto, 1982, 138 pp. 27. Polla, M., A Study of Nodal Regions in Strut-and-Tie Models, MASc thesis, Department of Civil Engineering, University of Toronto, 1992, 130 pp. 28. Barton, D. L. et al., Investigation of Strut-and-Tie Models for Dapped Beam Details, Research Report (Interim), Final Report, University of Texas at Austin, Center for Transportation Research, CTR-3-5-87/9-1127-1, FHWA/TX-92+1127-1, NTIS No. PB92-227735/HDM, May 1991, 205 pp. 29. Jirsa, J. O. et al., Experimental Studies of Nodes in Strut-and-Tie Models, Structural Concrete, IABSE Colloquium, Stuttgart, International Association for Bridge and Structural Engineering, Zrich, Mar. 1991, pp. 525-532. 30. Kuchma, D. A., and Tjhin, T. N., Design of Discontinuity Regions in Structural Concrete using a Computer-Based Strut-and-Tie Methodology, Proceedings of 81st TRB Annual Meeting, Washington, D.C., Jan. 2002.

594

ACI Structural Journal/September-October 2002

Das könnte Ihnen auch gefallen