You are on page 1of 47

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

net/publication/279039330

Evaluation of Existing Strut-and-Tie Methods and Recommended


Improvements

Article  in  Aci Structural Journal · November 2014


DOI: 10.14359/516869926

CITATIONS READS
13 245

5 authors, including:

Robin Tuchscherer
Northern Arizona University
20 PUBLICATIONS   86 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Robin Tuchscherer on 30 July 2019.

The user has requested enhancement of the downloaded file.


1 EVALUATION OF EXISTING STRUT AND TIE METHODS AND

2 RECOMMENDED IMPROVEMENTS

3 Robin G. Tuchscherer, David B. Birrcher, Christopher S. Williams, Dean J. Deschenes, and

4 Oguzhan Bayrak

5 Biography: ACI member Robin Tuchscherer is an Assistant Professor at Northern Arizona

6 University. He received his BS from the University of Wisconsin-Milwaukee; MS and PhD

7 from the University of Texas at Austin. He is a member of ACI Subcommittees 445-A, Strut

8 and Tie Modeling and 228-B, Visual Inspection; and an Associate Member of Committees

9 228, Nondestructive Testing of Concrete and 445, Shear and Torsion.

10 David Birrcher is a Bridge Engineer at International Bridge Technologies, Inc. in San Diego,

11 CA. He received his BS from Tulane University and his MS and PhD from the University of

12 Texas at Austin. His interests include improving the design and construction of reinforced

13 and prestressed concrete structures.

14 ACI member Christopher S. Williams is a PhD Candidate at the University of Texas at

15 Austin. He received his BS from Southern Illinois University Carbondale and his MS from

16 the University of Texas at Austin.

17 Dean J. Deschenes is a Research Associate at the Phil M. Ferguson Structural Engineering

18 Laboratory at the University of Texas at Austin. He received his BS from Lehigh University

19 and MS from the University of Texas at Austin.

20 Oguzhan Bayrak, FACI, is a Professor in the Department of Civil, Environmental, and

21 Architectural Engineering and holds the Charles Elmer Rowe Fellowship in Engineering at

22 the University of Texas at Austin, where he serves as Director of the Phil M. Ferguson

23 Structural Engineering Laboratory. He is a member of ACI Committees 341, Earthquake-

24 Resistant Concrete Bridges, and E803, Faculty Network Coordinating Committee, and Joint

25 ACI-ASCE Committees 441, Reinforced Concrete Columns, and 445, Shear and Torsion.

1
1 ABSTRACT

2 The objective of the study summarized in this paper was to develop simple, accurate and

3 conservative guidelines for the design of deep beams and discontinuity regions. In order to

4 accomplish this goal, the authors compiled a database of 868 deep beam shear tests from the

5 literature; and fabricated and tested 37 additional deep beam specimens. The cross-sectional

6 dimensions of the specimens that were tested included 36 x 48 in. (910 x 1220 mm); 21 x 75

7 in. (530 x 1910 mm); 21 x 42 in. (530 x 1070 mm); and 21 x 23 in. (530 x 580 mm). A

8 comprehensive analysis of the database resulted in the following conclusions: (i) deep beam

9 shear design provisions of ACI 318-11 and AASHTO LRFD (2010) are overly conservative;

10 and (ii) the accuracy of existing design provisions can be improved without compromising

11 their conservative nature. Accordingly, the authors propose recommendations to these

12 methods to form an improved strut-and-tie modeling procedure. The transparently derived

13 procedure is based on fundamental principles, current provisions, and experimental data. The

14 authors recommend adoption of the proposed modifications into the ACI 318 and AASHTO

15 LRFD provisions. An example problem illustrating the use of the proposed method is

16 provided in the Appendix.

17 Keywords: Strut-and-tie modeling, deep beam, shear, efficiency factor, triaxial confinement

18 BACKGROUND

19 Reinforced concrete members are typically designed to resist shear and flexural forces

20 based on Bernoulli’s beam theory, which states that strains vary linearly at a section or, that

21 “planar sections remain planar”. This assumption ignores the influence of shear

22 distortions/strains and is valid provided the shear span-to-depth ratio (a/d) is greater than 2 to

23 2.5. The ability of a Bernoulli or slender beam to resist loads is dependent on the geometric

24 and material properties of the cross-section. External loads and internal resistance are

25 equilibrated at a cross-section. On the other hand, deep beams (or discontinuity regions)

2
1 allow externally applied load to flow directly to the near support. Their resistance is less

2 dependent on the properties of the cross-section and more dependent on the configuration of

3 the structure at the discontinuities. Consequently, the mechanism of load transfer for a

4 slender beam is markedly different from that of a deep beam. Thus, the capacity of a deep

5 beam must be determined by an analytical method other than a sectional analysis. According

6 to ACI 318-111, deep beams may be designed either by taking into account the nonlinear

7 distribution of strains or by the strut-and-tie modeling procedure. Given the relative difficulty

8 of a nonlinear analysis, most practicing engineers prefer the strut-and-tie modeling procedure.

9 A strut-and-tie model (STM) idealizes the complex flow of stresses in a structural member by

10 aligning uniaxial truss elements with the resultants of the stress trajectories. Concrete struts

11 are aligned with the resultant of the compressive stress fields, and reinforcing steel ties are

12 aligned with the resultant of the tensile stress fields. The point of intersection of struts, ties, or

13 a combination of both forms a node. Struts, ties, and nodes are the three elements that shape a

14 STM and they must be proportioned to resist the applied forces. Based on the lower bound

15 theory of plasticity, the capacity of a STM is always less than or equal to the capacity of the

16 structure provided that: 1) the truss model is in equilibrium; 2) sufficient deformation

17 capacity exists to distribute forces within the assumed truss model; and 3) stresses applied to

18 elements do not exceed their yield capacity. The yield capacity of a strut or node is equal to

19 the effective compressive strength of concrete within the strut or at the interface between the

20 strut and node. The yield capacity of a tie is equal to the lesser of the tensile strength of the

21 steel reinforcement or force that causes a loss of anchorage.

22 The building code, ACI 318-111, and the bridge design specifications, AASHTO

23 LRFD (2010)2, adopted the use of strut-and-tie modeling in 2002 and 1994, respectively. A

24 detailed overview of the standard of practice in Europe is contained within a CEB-FIP (now

25 fib) report3, published in 1999. This document contains insight and commentary as to the

3
1 basis of the STM provisions in the current European Standard, EN 1992-1-1:20044 (i.e.

2 “Eurocode 2”).

3 RESEARCH SIGNIFICANCE

4 Variations exist between the strut-and-tie modeling procedures contained within ACI

5 318-11, AASHTO LRFD (2010), and fib (1999). In addition, due to the relative versatility of

6 the STM method, these provisions contain inconsistencies, ambiguities, or both. The purpose

7 of this paper is to evaluate current STM design provisions and establish a clear, transparent

8 and consistently defined methodology that practitioners can confidently employ for the

9 design of typical deep beams or discontinuity regions. The authors have calibrated and

10 substantiated the proposed procedure against a database of test results, theoretical principles,

11 and the current standard of practice.

12 DEEP BEAM DATABASE

13 The authors compiled a database of 868 deep beam shear tests (a/d ≤ 2.5) from

14 previous literature5-64. To supplement the data, the authors conducted 37 additional tests65-68.

15 The database containing all 905 tests is referred to as the “collection database.” The

16 collection database was filtered in two stages. In the first stage, test results that did not

17 contain the information necessary to perform a STM analysis were removed. The resulting

18 database is referred to as the “filtered database.” In the second stage, the authors removed

19 relatively small-scale and lightly-reinforced specimens. The authors filtered these tests so that

20 recommendations would be calibrated using specimens that are most representative of

21 transfer girders designed with STM in practice – i.e., reinforced and relatively large-scale. In

22 addition, in the authors’ experience, there are additional benefits to conducting large-scale

23 tests. Maintaining dimensional accuracy in concrete geometry, reinforcing bar details and test

24 setups in highly-scaled specimens is difficult. Realistically scaling down concrete mixture

25 designs and aggregate sizes for highly-scaled specimens is impractical. The database that

4
1 resulted from the second phase of filtering is referred to as the “evaluation database”. It

2 contains 179 deep beam shear tests and is deemed representative, accurate and

3 comprehensive.

4 Table 1 presents an overview of the database filtering process. Fig. 1 illustrates the

5 characteristics of the specimens within the evaluation database. Birrcher et al.69 provide

6 additional details for the data in the evaluation database.

7 SELECTION OF A STRUT-AND-TIE MODEL

8 The first step necessary to develop a STM procedure is to explicitly define the model.

9 The importance of this step cannot be over-emphasized as the predicted capacity of each

10 beam within the database is coupled to the configuration assumed for the model. In addition,

11 because the predicted capacity of a STM is dependent on the capacity of its weakest element,

12 the STM design procedures must be considered in their entirety. The authors evaluated the

13 capacity of each STM element for each beam simultaneously. In order to maintain

14 consistency with current standards of practice, the authors selected a single-panel truss model

15 with non-hydrostatic nodal zones to represent all of the deep beam shear tests in the

16 evaluation database. Fig. 2 illustrates the details of this model. The nodal regions were

17 proportioned in accordance with ACI 318-11, AASTHO LRFD (2010), and fib (1999). For

18 the reader’s convenience, the nodal proportions are illustrated in Fig. 3. Nodes are denoted by

19 the confluence of forces that abut them. For example, three struts bound a “CCC” node and

20 two struts and a tie bound a “CCT” node – where the “C” signifies a compressive force and

21 the “T” a tensile force.

22 A single-panel truss model was selected to maintain consistency with ACI 318-11 and

23 because a single strut is the primary mechanism of load transfer26, 69 for members with a/d

24 less than two. With that said, it is important to note that specimens with a/d as large as 2.5

25 were included in the analysis in order to assess the limits of the proposed procedure. A single

5
1 strut may not be a representative mechanism of shear transfer for beams with a/d of 2.5.

2 However, an overarching goal of this study was to evaluate the transition between deep beam

3 and sectional shear behavior.

4 Nodal regions of a STM are assumed to be either hydrostatic or non-hydrostatic. The

5 dimensions of a hydrostatic node would be proportioned such that stresses on all three faces

6 are equal. Shear stresses would not exist within a node that is in a hydrostatic state of stress.

7 However, the requirement of equal stresses on all faces of a node is rarely realized in practice,

8 as the resulting node dimensions are impractical. This phenomenon is illustrated in Fig. 4. As

9 a/d increases, the nodal proportions are increasingly inconsistent with the tie location or

10 depth of the flexural-compressive region. Thus, non-hydrostatic nodes were selected for the

11 beams evaluated for this study. This assumption is consistent with ACI 318-11, AASHTO

12 LRFD (2010), and fib (1999).

13 EVALUATION OF CURRENT DESIGN PROVISIONS

14 Based on the truss model illustrated in Fig. 2 and the experimental results of the 179

15 specimens in the evaluation database, a statistical comparison was made between the ACI

16 318-11, AAASHTO LRFD (2010), and fib (1999) strut-and-tie modeling provisions. Table 2

17 includes a list of the allowable stress or maximum concrete efficiency stipulated by each of

18 these provisions. Stresses are simultaneously checked at the following seven locations for all

19 beams in the database: 1) back face of the CCC node; 2) back face of the CCT node; 3)

20 bearing face of the CCC node; 4) bearing face of the CCT node; 5) strut-to-node interface at

21 the CCC node; 6) strut-to-node interface of the CCT node and; 7) within the main

22 longitudinal tie reinforcement. These seven stress checks account for all possible failure

23 modes with two exceptions: 1) proper anchorage of the tie must be provided and; 2) the beam

24 must contain a minimum amount of web reinforcement to equilibrate the transverse spreading

25 of compression and prevent premature splitting of the bottle-shaped strut. All of the

6
1 specimens in the evaluation database contained adequate anchorage and a minimum amount

2 of web reinforcement to avoid these premature failure modes. The ratio of experimental to

3 calculated capacity was determined for the beams in the evaluation database. Fig. 5 presents

4 histograms of the findings.

5 Upon observation of Fig. 5, the estimated capacities calculated per the fib (1999)

6 provisions were significantly more accurate than the ACI 318-11 and AASHTO LRFD

7 (2010) provisions (COV of 0.27 versus 0.59 and 0.68, respectively). This is the result of

8 many contributing factors, the most significant being the fact that fib (1999) allows the

9 effective strength of a node to be increased when triaxial confinement is present. To illustrate,

10 by ignoring triaxial confinement, ACI 318-11 and AASHTO LRFD (2010) conservatively

11 estimate the capacity of two specimens to be approximately ten times less than their actual

12 strength (experimental to predicted ratios of 9.80 and 11.77, respectively). Other significant

13 factors include the fact that: fib (1999) allows bonding stresses at the back face of a CCT

14 node to be ignored; and the fib (1999) efficiency factors increase at a diminishing rate as the

15 compressive strength of concrete increases.

16 The AASHTO LRFD (2010) provisions are less accurate than the ACI 318-11

17 provisions due their respective CCT efficiency factors. The AASHTO LRFD (2010)

18 provisions specify a CCT efficiency factor that decreases with increasing a/d. Likely, the

19 factor has been calibrated with hydrostatic nodes (that increase in area with increasing a/d).

20 When the AASHTO LRFD (2010) CCT efficiency factors are used in combination with non-

21 hydrostatic nodes, the procedure results in an overly conservative estimation of strength.

22 DEVELOPMENT OF PROPOSED DESIGN PROCEDURE

23 Again, the objective of this study was to improve upon current design procedures.

24 According to MacGregor70, a “design procedure” should satisfy the following four criteria: 1)

25 simplicity in application, 2) compatibility with tests of discontinuity regions, 3) compatibility

7
1 with other sections of the code, and 4) compatibility with other codes or design

2 recommendations. In keeping with these precepts, the authors developed an STM design

3 procedure that is consistent with ACI 318-11 and AASHTO LRFD (2010) but also contains

4 key attributes of the fib (1999) provisions.

5 The following recommendations are calibrated in a clear and comprehensive manner.

6 The resulting six items represent the important characteristics of the strut and tie modeling

7 procedure proposed in this paper.

8 1. The allowable stress of all faces of a node may be increased when triaxial

9 confinement is present. Tuchscherer et al.65 tested ten deep beam specimens with

10 triaxially confined nodal regions. The authors concluded that increasing the capacity

11 of a triaxially confined node resulted in more accurate STM calculations without

12 adversely affecting the level of conservatism. This conclusion is similar with fib

13 (1999). ACI 318-11 and AASHTO LRFD (2010) allow the bearing strength of

14 concrete to be increased by the factor 2 when the loaded area is surrounded

15 by concrete on all sides. In order to maintain consistency with these provisions, a

16 corresponding triaxial confinement factor is proposed for the design of STM nodal

17 regions. That is, the allowable capacity of concrete on all surfaces of a triaxially

18 confined node may be increased by the factor 2.

19 2. If tie reinforcement is properly anchored, then crushing of concrete at the node

20 face normal to the tie is unlikely to occur. The authors69 and previous researchers71-
73
21 have reached similar conclusions and this philosophy is consistent with fib (1999).

22 In a separate paper, Tuchscherer et al.74 examined the implications of ignoring the

23 stress applied to the back face of the CCT node. Subsequently, the authors showed

8
1 that the elimination of a stress check at the back face of a CCT node did not have a

2 significant effect on the conservatism or accuracy of STM predictions. The utility of

3 performing stress checks under a “conventionally assumed” concentrated force at the

4 back face of the CCT nodes was unsubstantiated74.

5 3. The effective concrete stress at the back face and bearing face of a CCC node

6 should be taken equal to 0.85f´c. The normalized stress at the back and bearing face

7 of the CCC node is plotted for all beams where said nodal boundaries control the

8 design strength (Fig. 6). Admittedly, there are a sparse number of tests that are

9 controlled by the CCC bearing or back face. Nevertheless, based on the results

10 obtained from the database, a constant efficiency of 0.85 is a conservative value; is

11 consistent with ACI 318-11 and AASHTO LRFD (2010) provisions; and is in the

12 interest of having the simplest code provision that captures the trend (or lack thereof)

13 of the data presented in Fig. 6.

14 4. The effective concrete stress at the bearing face of a CCT node should be taken

15 equal to 0.70f'c. The allowable compressive strength of concrete in a CCT nodal

16 region is assumed to be less than in a CCC nodal region. This philosophy, adopted by

17 ACI 318-11 and AASHTO LRFD (2010), is based on the observation75 that the

18 accumulation of transverse tensile cracks caused by the straining of the tie reduces the

19 compressive strength of concrete. To maintain consistency with these provisions, a

20 similar efficiency factor was examined for the proposed method. The implication of

21 this proposal was assessed by plotting the stress at the CCT bearing face for all beams

22 where said boundary controlled the design strength (Fig. 7). While there are a sparse

23 number of beams whose capacity was controlled by the CCT bearing face, a constant

24 efficiency of 0.70 is a lower bound value and is consistent with ACI 318-11 and

25 AASHTO LRFD (2010).

9
1 5. The effective concrete stress at the CCC strut-to-node interface should be the

2 same as the effective stress at the CCT strut-to-node interface. ACI 318-11

3 specifies an identical efficiency at both the CCC and CCT strut-to-node interface.

4 AASHTO LRFD (2010) and fib (1999) specify a lower efficiency at the CCT strut-to-

5 node interface. These philosophies are examined against the trends in the database.

6 The capacity of all of the beams was calculated according to fib (1999), ACI 318-11,

7 and AASHTO LRFD (2010). If the stress at the CCT or CCC strut-to-node interface

8 controlled, the actual stress applied to this surface was noted. Upon observation of the

9 results, the capacity of the CCC interface did not control for any of the beams per fib

10 (1999) with the exception of one instance. Nor did the CCC interface control per

11 AASHTO LRFD (2010) with the exception of fourteen “panel tests” (a/d = 0). The

12 results for beams whose capacities were estimated per ACI 318-11 are presented in

13 Fig. 8. As can be seen, the data is equally scattered and has a similar lower bound.

14 Because there is no evidence to support the use of separate efficiencies, and to

15 maintain consistency with ACI 318-11, the authors recommend an identical efficiency

16 factor for both the CCC and CCT interfaces.

17 6. As the compressive strength of concrete increases, the effective stress at the

18 strut-to-node interface increases at a diminishing rate. High-strength concrete is

19 typically that which has a compressive strength greater than 8000 psi (55 MPa).

20 Generally, the cement paste of high-strength concrete is stronger than the aggregate

21 and the quality of aggregate heavily influences its ability to transfer force by

22 aggregate interlock. Previous researchers11, 76-78 and fib (1999) similarly recognize that

23 an increase in the compressive strength of concrete is associated with a comparable

24 reduction in its efficiency to transfer shear, tension, or both. As a result, the authors

25 propose an efficiency factor at the strut-to-node interfaces that diminishes similar to

10
1 the factor recommended by fib (1999) but is also consistent with ACI 318-11 and

2 AASHTO LRFD (2010). This efficiency factor is derived as follows:

3 Currently, ACI 318-11 STM provisions require the efficiency factor at the

4 strut-to-node interface to be a constant value equal to 0.64 (Table 2). This value is

5 calculated by taking 0.85 multiplied by 0.75. In the interest of simplifying the existing

6 provision and maintaining consistency, the authors propose a maximum efficiency of

7 0.65. The highest compressive strength in the evaluation database is approximately

8 17,400 psi (120 MPa). The fib (1999) strut-to-node efficiency factor is equal to 0.48

9 when the compressive strength of concrete is 17,400 psi (120 MPa). Given that there

10 is not much data available in the high-strength range, and in the interest of simplicity

11 and consistency, the authors propose a minimum efficiency 0.45 for compressive

12 strengths greater than 8000 psi (55 MPa). Finally, the authors propose a linearly

13 decreasing efficiency factor for compressive strengths between 4000 and 8000 psi (28

14 and 55 MPa). To summarize, the efficiency factor, ν, at both the CCC and CCT strut-

15 to-node interface is listed in Eq. 1:

16 0.45 0.85 20 0.65 (1)

17 0.45 0.85 138 0.65 [S.I.]

18 The implication of this proposal is assessed against the evaluation database.

19 The normalized stress at the CCC and CCT node-to-strut interface is plotted for all

20 beams where said boundary controls the design strength (Fig. 9). As shown, based on

21 the results obtained from the database, the efficiency factor proposed for the strut-to-

22 node interface (Eq. 1) is an appropriate expression.

23 SYNOPSIS OF PROPOSED DESIGN PROCEDURE

24 The above recommendations are summarized as follows.

11
1 Fig. 3 illustrates the proportions of the CCC and CCT nodal regions. Eq. 2 calculates

2 the nominal strength of these nodal zones, Fn.

3 ∙ (2)

4 Where,

5 fce = effective compressive strength of concrete in nodal zone

6 Anz = cross-sectional area of the face of the nodal zone

7 Eq. 3 determines the effective compressive strength of the face of a nodal zone, fce.

8 ∙ ∙ (3)

9 Where,

10 m= 2 triaxial confinement modification factor

11 A1 = loaded area

12 A2 = area of the lower base of the largest frustum of a pyramid, cone, or tapered

13 wedge contained wholly within the support and having for its upper base the

14 loaded area and having a side slope of 1 vertical to 2 horizontal (refer to

15 Figure R10.14 of ACI 318-11)

16 fc´ = specified compressive strength of concrete

17 ν= concrete efficiency factor

18 = 0.85 for bearing and back face of CCC node

19 = 0.70 for bearing face of CCT node

20 = 0.45 0.85 20 0.65 for CCC and CCT strut-to-node interface

21 when minimum transverse reinforcement

22 is provided

12
0.45 0.85 138 0.65

1 The efficiency factor proposed at the CCC and CCT strut-to-node interface assumes

2 that the member contains a minimum amount of web reinforcement. In a companion paper,

3 Birrcher et al.67 recommend a minimum web reinforcement ratio of 0.002 in both the

4 horizontal and vertical directions to prevent premature loss of equilibrium, and a ratio of

5 0.003 in both directions to limit diagonal cracks to less than 0.016 in. (0.4 mm) under the

6 application of service loads.

7 EVALUATION OF PROPOSED DESIGN PROCEDURE

8 Fig. 5 presents the results of a statistical analysis of the proposed design procedure

9 alongside the ACI 318-11, AASHTO LRFD (2010), and fib (1999) STM provisions. As

10 shown, the proposed strut-and-tie modeling design procedure is a significant improvement

11 over the current ACI 318-11 and AASHTO LRFD (2010) procedures. The fib (1999)

12 procedure is slightly more accurate than the proposed procedure (COV of 0.27 versus 0.29,

13 respectively); however, this trend is attributed to the difference in triaxial confinement factors.

14 The proposed procedure limits the triaxial confinement modification factor to a value of two

15 whereas fib (1999) limits the value to four. For the basis of comparison, if the proposed

16 method were to limit the confinement factor to four, then it would perform slightly better than

17 the fib (1999) method (COV of 0.24 versus 0.27, respectively). Nonetheless, the authors

18 propose to limit the triaxial confinement factor to two to maintain consistency with ACI 318-

19 11 and AASHTO LRFD (2010). The merits of increasing the factor to a value of four were

20 not an objective of this study.

21 All of the design provisions were evaluated against the experimental results of the

22 specimens in the evaluation database. All of these specimens contain a minimum amount of

23 web reinforcement equal to approximately 0.14% of the cross-sectional area in both the

13
1 horizontal and vertical direction (the minimum ratio of reinforcement perpendicular to the

2 diagonal strut, ﬩, for all beams in the evaluation database is greater than 0.1% of the cross-

3 sectional area). This amount is significantly less than the minimum amount required by ACI

4 318-11, AASHTO LRFD (2010), and fib (1999). The AASHTO LRFD (2010) and fib (1999)

5 provisions require a minimum amount of crack control reinforcement equal to 0.3% and 0.2%

6 of the cross-sectional area in both the vertical and horizontal direction, respectively. ACI

7 318-11 STM provisions do not require minimum web reinforcement; however, efficiency

8 factors are reduced when ﬩ is less than 0.3%.

9 To assess whether earlier conclusions with regard to accuracy and conservatism

10 remain valid for varying amounts of web reinforcement, each of the design provisions are

11 examined against trends in the database. However, the difficulty of evaluating beams with,

12 for example, the AASHTO LRFD (2010) minimum reinforcement requirement of 0.3% in

13 each direction is that only a sparse number of beams in the literature meet this requirement.

14 Specifically, only 45 and 26 beams meet the fib (1999) and AASHTO LRFD (2010)

15 minimum requirements, respectively. As such, significant conclusions should not be inferred

16 from statistical comparisons of these small datasets. Nonetheless, for the purpose of

17 comparison, the design provisions are examined against only those beams that contain the

18 minimum reinforcement requirements specified above. A summary of the findings is

19 presented in Fig. 10.

20 As expected, a reduction of the number of beams in the dataset results in an increase

21 in the COV. In addition, as the amount of web reinforcement increases, the number of

22 unconservative estimations of strength decreases. While a reduction in data causes a rise in

23 COV, the relative magnitude of the COV value between each provision is not significantly

24 altered by quanitty of transverse reinforcement. As a result, it is concluded that the 179

14
1 specimens in the evaluation database that contain a strut reinforcement ratio, ﬩, greater than

2 or equal to 0.1% provide a valid basis of comparison between the different provisions.

3 SUMMARY AND CONCLUSIONS

4 The authors developed a STM procedure for the design of deep beams. In developing

5 this procedure, it was necessary to explicitly define the configuration of the elements of a

6 STM. The proposed procedure is based on a simple single-panel truss with non-hydrostatic

7 nodal regions. The authors anticipate that practicing engineers will be able to use this model

8 as a basis for determining the configuration of many other discontinuity regions encountered

9 in practice. With that said, a limitation to note is that the proposed procedure was calibrated

10 only considering specimens with a width, bw, greater than 4.5 in. (110 mm), cross-section,

11 bwd, greater than 100 in2 (645 cm2), and web reinforcement ratio greater than 0.1% (Table 1).

12 An important aspect of the recommended STM procedure is that the authors

13 comprehensively derived it by considering all of the stress checks necessary to design the

14 members within the database. The splitting of the strut was indirectly accounted for by only

15 considering those beams that contained a minimum amount of transverse reinforcement (ρ﬩ ≥

16 0.1%), and the primary tie reinforcement was adequately anchored. Thus, the derivation of

17 the proposed procedure considers every facet of an STM design.

18 In conclusion, the proposed STM procedure is simpler and more accurate than the

19 ACI 318-11 and AASHTO LRFD (2010) provisions. The procedure is based on established

20 principles of a strut-and-tie design; on experimental tests of deep beams; and on the

21 provisions within ACI 318-11, AASHTO LRFD (2010), and fib (1999). The procedure is

22 practical and has been derived in a clear and comprehensive manner.

23 An example problem illustrating the use of the proposed STM procedure for the

24 design of a reinforced concrete transfer girder is provided in the Appendix.

15
1 ACKNOWLEDGMENTS

2 The authors wish to express their gratitude and sincere appreciation to the Texas

3 Department of Transportation (TxDOT) for providing the funding to conduct this research

4 study. The recommendations of project director Dean Van Landuyt (Bridge Division), John

5 Vogel (Houston District), and David Hohmann (Bridge Division) are deeply appreciated.

6 Also, the authors would like to thank Matt Huizinga for his technical contributions and

7 assistance with the experimental portion of the project. Finally, the recommendations of

8 James Jirsa, Sharon Wood, and John Breen improved the quality of this research and are

9 greatly valued.

10

16
1 NOTATION

2 a= shear span
3 a/d = shear span-to-depth ratio measured center of span to center of support
4 A1 = loaded area
5 A2 = area of the lower base of the largest frustum of a pyramid, cone, or tapered wedge
6 contained wholly within the support and having for its upper base the loaded area and
7 having side slopes of 1 vertical to 2 horizontal (refer to Figure R10.14 of ACI 318-11)
8 Anz = cross-sectional area of the face of a nodal zone
9 As = area of tension reinforcement
10 A’s = area of compression reinforcement
11 bw = width of beam web
12 c= distance from extreme compression fiber to neutral axis
13 d= depth of member taken as the distance from extreme compression fiber to centroid of
14 longitudinal tension reinforcement
15 f’c = specified compressive strength of concrete
16 fce = effective compressive strength of concrete in nodal zone
17 Fn = nominal strength of a nodal zone
18 fs = stress in tension reinforcement
19 f’s = stress in compression reinforcement
20 fy = specified yield strength of tensile reinforcement
21 h= overall height of member
22 ha = height of the back face of a CCT node taken as twice the distance from extreme
23 tension fiber to centroid of longitudinal tension reinforcement
24 hs = height of the back face of a CCC node taken as the height of the equivalent stress
25 block of the flexural compression zone
26 ll = length of the bearing plate at the CCC node
27 ln = span (from MacGregor5)
28 ls = length of the bearing plate at the CCT node

29 m= triaxial confinement modification factor = 2

30 P= axial load
31 α= proportion of applied load that is resisted by the near support

17
1 β1 = factor for proportioning the depth of the equivalent stress block in the flexural
2 compression region
3 ε1 = principle tensile strain in cracked concrete
4 h = ratio of horizontal shear reinforcement to gross concrete area of vertical section
5 v = ratio of vertical shear reinforcement to gross concrete area of horizontal section
6 = ratio of the shear reinforcement perpendicular to the plane of the strut
7 θ= angle of strut measured from the horizontal axis
8 ν= efficiency factor, concrete effectiveness factor

18
1 REFERENCES

2 1. ACI Committee 318, Building Code Requirements for Structural Concrete (ACI 318-11)
3 and Commentary (ACI 318R-11), American Concrete Institute, Farmington Hills, MI,
4 2011, 503 pp.
5 2. AASHTO 2010, LRFD Bridge Design Specifications, 5th Edition, with 2010 Interim
6 Revisions, American Association of State Highway and Transportation Officials,
7 Washington D. C., 2010, 1822 pp.
8 3. FIP Report, Practical Design of Structural Concrete, Federation Internationale de la
9 Precontrainte, CEB-FIP (fib), Federation Internationale du Beton, Lausanne,
10 Switzerland, September 1999, 114 pp.
11 4. Eurocode 2: Design of concrete structures – Part 1-1: General rules and rules for
12 Buildings, English Version, European Committee for Standardization, CEN, Brussels,
13 Belgium, April 2004, 225 pp.
14 5. Ahmad, S. H., and Lue, D. M., “Flexure-Shear Interaction of Reinforced High-Strength
15 Concrete Beams,” ACI Journal, No. 84, July-August 1987, pp. 330-341.
16 6. Alcocer, S. M., and Uribe, C. M., “Monolithic and Cyclic Behavior of Deep Beams
17 Designed Using Strut-and-Tie Models,” ACI Journal, No. 105, May-June 2008, pp.
18 327-337.
19 7. Angelakos, D., The Influence of Concrete Strength and Longitudinal Reinforcement
20 Ratio on the Shear Strength of Large-Size Reinforced Concrete Beams With, and
21 Without, Transverse Reinforcement, Master’s Thesis, 1999, Department of Civil
22 Engineering, University of Toronto, 195 pp.
23 8. Angelakos, D.; Bentz, E. C; and Collins, M. P., “Effect of Concrete Strength and
24 Minimum Stirrups on Shear Strength of Large Members,” ACI Journal, No. 98, May-
25 June 2001, pp. 290-300.
26 9. Bažant, Z. P., and Kazemi, M. T., “Size Effect of Diagonal Shear Failure of Beams
27 without Stirrups,” ACI Journal, No. 88, May-June 1991, pp. 268-276.
28 10. Bresler, B., and Scordelis, A. C., “Shear Strength of Reinforced Concrete Beams,” ACI
29 Journal, No. 60, January 1963, pp. 51-74.
30 11. Brown, M. D.; Sankovich, C. L.; Bayrak, O.; Jirsa, J. O.; Breen, J. E.; and Wood, S. L.,
31 Design for Shear in Reinforced Concrete Using Strut-and-Tie Models, Report No. 0-
32 4371-2, Center for Transportation Research, University of Texas at Austin, Austin,
33 Texas, Apr. 2006.
34 12. Brown, M. D.; Sankovich, C. L.; Bayrak, O.; Jirsa, J. O., “Behavior and Efficiency of
35 Bottle-Shaped Struts,” ACI Journal, No. 103, May-June 2006, pp. 348-355.
36 13. Cao, S., Size Effect and the Influence of Longitudinal Reinforcement on the Shear
37 Response of Large Reinforced Concrete Members, Master’s Thesis, University of
38 Toronto, Toronto, Ontario, Canada, 2001, 195 pp.
39 14. Chang, T. S., and Kesler, C. E., “Static and Fatigue Strength in Shear of Beams with
40 Tensile Reinforcement,” ACI Journal, No. 54, June 1958, pp. 1033-1057.
41 15. Clark, A. P., “Diagonal Tension in Reinforced Concrete Beams,” ACI Journal, No. 48,
42 October 1951, pp. 145-156.

19
1 16. de Cossio, R. D., and Siess, C. P., “Behavior and Strength in Shear of Beams and Frames
2 without Web Reinforcement,” ACI Journal, No. 56, February 1960, pp. 695-735.
3 17. de Paiva, H. A. R., and Siess, C. P., ”Strength and Behavior of Deep Beams in Shear,”
4 Journal of the Structural Division, ASCE Proceedings, ST 5, October 1965, pp. 19-41.
5 18. Ferguson, P. M., “Some Implications of Recent Diagonal Tension Tests,” ACI Journal,
6 No. 53, August 1956, pp. 157-172.
7 19. Foster, S. J., and Gilbert, R. I., “Experimental Studies on High-Strength Concrete Deep
8 Beams,” ACI Journal, No. 95, July-August 1998, pp. 382-390.
9 20. Furuuchi, H.; Takahashi, Y.; Ueda, T.; and Kakuta, Y., “Effective Width for Shear
10 Failure of RC Deep Slabs,” Transaction of the Japan Concrete Institute, Vol. 20,
11 1998, pp. 209-216.
12 21. Ghoneim, M., “Shear Strength of High-Strength Concrete Deep Beams,” Journal of
13 Engineering and Applied Science, Vol. 48, No. 4, August 2001, pp. 675-693.
14 22. Hara, T., The Shear Strength of Reinforced Concrete Deep Beams, Transaction of the
15 Japan Concrete Institute, Vol. 6, 1985, pp. 395-402.
16 23. Hassan, T. K.; Seliem, H. M; Dwairi, H.; Rizkalla, S. H.; Zia, P., “Shear Behavior of
17 Large Concrete Beams Reinforced with High-Strength Steel,” ACI Journal, No. 105,
18 March-April 2008, pp. 173-179.
19 24. Hsuing, W. and Frantz, G. C., “Transverse Stirrup Spacing in R/C Beam,” ASCE Journal
20 of Structural Engineering, Vol. 11, No. 2, February 1985, pp. 353-362
21 25. Johnson, M. K., and Ramirez, J. A., “Minimum Shear Reinforcement in Beams with
22 Higher Strength Concrete,” ACI Journal, No. 86, July-August 1989, pp. 376-382.
23 26. edited by Kani, M. W.; Huggins, M. W.; and Wittkopp, R. R., Kani on Shear in
24 Reinforced Concrete, University of Toronto Press, Toronto, 1979, 225 pp.
25 27. Kong, P. Y. L., and Rangan, B. V., “Shear Strength of High-Performance Concrete
26 Beams,” ACI Journal, No. 95, November-December 1998, pp. 677-688.
27 28. Kong, F.; Robins, P. J.; and Cole D. F., “Web Reinforcement Effects on Deep Beams,”
28 ACI Journal, No. 67, December 1970, pp. 1010-1018.
29 29. Krefeld, W. J., and Thurston, C. W., “Studies of the Shear and Diagonal Tension
30 Strength of Simply Supported Reinforced Concrete Beams,” ACI Journal, No. 63,
31 April 1966, pp. 451-476.
32 30. Krefeld, W. J., and Thurston, C. W., “Contribution of Longitudinal Steel to Shear
33 Resistance of Reinforced Concrete Beams,” ACI Journal, No. 63, March 1966, pp.
34 325-344.
35 31. Leonhardt, F. and Walther, R., translation by Amerongen, C. V., “The Stuttgardt Shear
36 Tests, 1961”, from Beton und Stahlbeton, Vol. 56, No. 12, 1961 and Vol. 57, No. 2, 3,
37 6, 7, and 8, 1962, Translation No. 111, Cement and Concrete Association, London,
38 1964, 138 pp.
39 32. Manuel, R. F.; Slight, B. W.; and Suter, G. T., “Deep Beam Behavior Affected by
40 Length and Shear Span Variables,” ACI Journal, No. 68, December 1971, pp. 954-
41 958.

20
1 33. Matsuo, M.; Lertsrisakulrat, T.; Yanagawa, A.; and Niwa, J., “Shear Behavior of RC
2 Deep Beams with Stirrups,” Transaction of the Japan Concrete Institute, Vol. 23,
3 2002, pp. 385-390.
4 34. Moody, K. G.; Viest, I. M.; Elstner, R. C.; and Hognestad, E., “Shear Strength of
5 Reinforced Concrete Beams, Part 1 – Tests of Simple Beams,” ACI Journal, No. 51,
6 December 1954, pp. 317-332.
7 35. Morrow, J., and Viest, I. M., “Shear Strength of Reinforced Concrete Frame Members
8 without Web Reinforcement,” ACI Journal, No. 53, March 1957, pp. 833-869.
9 36. Oh, J., and Shin, S., “Shear Strength of Reinforced High-Strength Concrete Deep
10 Beams,” ACI Journal, No. 98, March-April 2001, pp. 164-173.
11 37. Ozcebe, G.; Ersoy, U.; and Tankut, T., “Evaluation of Minimum Shear Reinforcement
12 Requirements for Higher Strength Concrete,” ACI Journal, No. 95, May-June 1999,
13 pp. 361-369.
14 38. Quintero-Febres, C. G.; Montesinos, G. P.; and Wight, J. K., “Strength of Struts in Deep
15 Concrete Members Designed Using Strut-and-Tie Method,” ACI Journal, No. 103,
16 July-August 2006, pp. 577-586.
17 39. Rajagopalan, K. S., and Ferguson, P. M., “Exploratory Shear Tests Emphasizing
18 Percentage of Longitudinal Steel,” ACI Journal, No. 65, August 1968, pp. 634-638.
19 40. Ramakrishnan, V. and Ananthanarayana, Y., “Ultimate Strength of Deep Beams in
20 Shear,” ACI Journal, No. 65, February 1968, pp. 87-98.
21 41. Rigotti, M., Diagonal Cracking in Reinforced Concrete Deep Beams – An Experimental
22 Investigation, PhD Dissertation, Concordia University, Montreal, Quebec, Canada,
23 November 2002, 235 pp.
24 42. Rogowsky, D. M.; MacGregor, J. G.; and Ong, S. Y., “Tests of Reinforced Concrete
25 Deep Beams,” ACI Journal, No. 83, July-August 1986, pp. 614-623.
26 43. Roller, J. J., and Russell, H. G., “Shear Strength of High-Strength Concrete Beams with
27 Web Reinforcement,” ACI Journal, No. 87, March-April 1990, pp. 191-198.
28 44. Sarsam, K. F., and Al-Musawi, J. M. S., “Shear Design of High and Normal Strength
29 Concrete Beams with Web Reinforcement,” ACI Journal, No. 89, November-
30 December 1992, pp. 658-664.
31 45. Shin, S.; Lee, K.; Moon, J.; and Ghosh, S. K., “Shear Strength of Reinforced High-
32 Strength Concrete Beams with Shear Span-to-Depth Ratios between 1.5 and 2.5,” ACI
33 Journal, No. 96, July-August 1999, pp. 549-557.
34 46. Shioya, T. S., Shear Properties of Large Reinforced Concrete Member, Special Report
35 of Institute of Technology, Shimizu Corporation, No. 25, February 1989, 213 pp.
36 47. Smith, K. N. and Vantsiotis, A. S., “Shear Strength of Deep Beams,” ACI Journal, No.
37 79, May-June 1982, pp. 201-213.
38 48. Stanik, B. A. P., The Influence of Concrete Strength, Distribution of Longitudinal
39 Reinforcement, Amount of Transverse Reinforcement and Member Size on Shear
40 Strength of Reinforced Concrete Members, Master’s Thesis, 1998, Department of
41 Civil Engineering, University of Toronto, 369 pp.

21
1 49. Subedi, N. K.; Vardy, A. E.; and Kubota, N., “Reinforced Concrete Deep Beams – Some
2 Test Results,” Magazine of Concrete Research, Vol. 38, No. 137, December 1986, pp.
3 206-219.
4 50. Tan, K.; Kong, F.; Teng, S.; and Guan, L.,”High-Strength Concrete Deep Beams with
5 Effective Span and Shear Span Variations,” ACI Journal, No. 92, July-August 1995,
6 pp. 1-11.
7 51. Tan, K.; Kong, F.; Teng, S.; and Weng, L., ”Effect of Web Reinforcement on High-
8 Strength Concrete Deep Beams,” ACI Journal, No. 94, September-October 1997, pp.
9 572-582.
10 52. Tan, K. H., and Lu, H. Y., “Shear Behavior of Large Reinforced Concrete Deep Beams
11 and Code Comparisons,” ACI Journal, No. 96, September-October 1999, pp. 836-
12 846.
13 53. Tan, K.; Teng, S.; Kong, F.; Lu, H., “Main Tension Steel in High Strength Concrete
14 Deep and Short Beams,” ACI Journal, No. 94, November-December 1997, pp. 752-
15 768.
16 54. Tanimura, Y., and Sato, T., “Evaluation of Shear Strength of Deep Beams with
17 Stirrups,” Quarterly Report of the Railway Technical Research Institute, Vol. 46, No.
18 1, February 2005, pp. 53-58.
19 55. Uzel, A., Shear Design of Large Footings, Ph. D. Dissertation, University of Toronto,
20 Toronto, Ontario, Canada, 2003, 404 pp.
21 56. Van Den Berg, F. J., “Shear Strength of Reinforced Concrete Beams without Web
22 Reinforcement,” ACI Journal, No. 59, November 1962, pp. 1587-1600.
23 57. Vecchio, F. J., “Analysis of Shear-Critical Reinforced Concrete Beams,” ACI Journal,
24 No. 97, January-February 2000, pp. 102-110.
25 58. Walraven, J., and Lehwalter, N., “Size Effects in Short Beams Loaded in Shear,” ACI
26 Journal, No. 91, September-October 1994, pp. 585-593.
27 59. Watstein, D., and Mathey, R. G., “Strains in Beams having Diagonal Cracks,” ACI
28 Journal, No. 55, December 1958, pp. 717-728.
29 60. Xie, Y.; Ahmad, S. H.; Yu, T.; Hino, S.; and Chung, W., “Shear Ductility of Reinforced
30 Concrete Beams of Normal and High-Strength Concrete,” ACI Journal, No. 91,
31 March-April 1954, pp. 140-149.
32 61. Yang, K.; Chung, H.; Lee, E.; and Eun, H., “Shear Characteristics of High-Strength
33 Concrete Deep Beams without Shear Reinforcements,” Engineering Structures, No.
34 25, April 2003, pp. 1343-1352.
35 62. Yoon, Y.; Cool, W. D.; and Mitchell, D., “Minimum Shear Reinforcement in Normal,
36 Medium, and High-Strength Concrete Beams,” ACI Journal, No. 93, September-
37 October 1996, pp. 1-9.
38 63. Yoshida, Y., Shear Reinforcement for large Lightly Reinforced Concrete Members,
39 Master’s Thesis, University of Toronto, Toronto, Ontario, Canada, 2000, 162 pp.
40 64. Zhang, N., and Tan, K., “Size Effect in RC Deep Beams: Experimental Investigation and
41 STM Verification,” Engineering Structures, No. 29, October 2007, pp. 3241-3254.

22
1 65. Tuchscherer, R. G.; Birrcher, D. B.; Huizinga, M. R.; and Bayrak, O., “Confinement of
2 Deep Beam Nodal Regions,” ACI Structural Journal, V. 107, No. 6, Nov.-Dec. 2010,
3 pp. 709-717.
4 66. Tuchscherer, R. G.; Birrcher, D. B.; Huizinga, M. R.; and Bayrak, O., “Distribution of
5 Stirrups across Web of Deep Beams,” ACI Structural Journal, V. 108, No. 1, Jan.-
6 Feb. 2011, pp. 108-115.
7 67. Birrcher, D. B.; Tuchscherer R. G.; Huizinga, M. R.; and Bayrak, O., “Minimum Web
8 Reinforcement in Deep Beams,” ACI Structural Journal, V. 110, No. 2, Mar.-Apr.
9 2013, pp. 297-306.
10 68. Birrcher, D. B.; Tuchscherer R. G.; Huizinga, M. R.; and Bayrak, O., “Depth Effect in
11 Deep Beams,” ACI Structural Journal, in press.
12 69. Birrcher, D. B.; Tuchscherer, R. G.; Huizinga, M. R.; Bayrak, O.; Wood, S. L.; Jirsa, J.
13 O.; Strength and Serviceability Design of Reinforced Concrete Deep Beams, Report
14 No. 0-5253-1, Center for Transportation Research, University of Texas at Austin,
15 Austin, TX, April 2009, 376 pp.
16 70. MacGregor, J. D., “Part 2: Derivation of strut-and-tie models for the 2002 ACI Code,”
17 ACI Special Publication, Examples for the Design of Structural Concrete with Strut
18 and Tie Models, SP-208, Sixth Printing, Aug. 2008, pp. 8-40.
19 71. Schlaich, J.; Schäfer, K.; and Jennewein, M., “Toward a Consistent Design of Structural
20 Concrete,” PCI Journal, Vol. 32, No. 3, Precast/Prestressed Concrete Institute,
21 Chicago, IL, May-June 1987, pp. 74-150.
22 72. Thompson, M. K.; Young, M. J.; Jirsa, J. O.; Breen, J. E.; and Klingner, R. E.,
23 Anchorage of Headed Reinforcement in CCT Nodes, Research Report 1855-2, Center
24 for Transportation Research, University of Texas at Austin, Austin, TX, June 2003,
25 146 pp.
26 73. Larson, N.; Gómez, E.; Garber, D.; Bayrak, O.; and Ghannoum, W., Strength and
27 Serviceability Design of Reinforced Concrete Inverted-T Beams, Report No. 6416-1,
28 Center for Transportation Research, University of Texas at Austin, Austin, Texas,
29 June 2013, 234 pp.
30 74. Tuchscherer, R.; Birrcher, D.; and Bayrak, O., “Experimental Examination of ACI 318
31 Strut and Tie Modeling Provisions,” ACI Special Publication, SP-296-7, Symposium
32 Honoring James O. Jirsa, sponsored by Joint ACI-ASCE Committees 352, 369, 374,
33 408, and 445, 2014, pp. 1-20.
34 75. Vecchio, F. J. and Collins, M. J., “The Modified Compression Field Theory for
35 Reinforced Concrete Element Subjected to Shear,” ACI Structural Journal, Vol. 83,
36 No. 2, March-April 1986, pp.219-231.
37 76. Bergmeister, K.; Breen, J. E.; Jirsa, J. O.; and Kreger, M. E., Detailing for Structural
38 Concrete, Report No. 1127-3F, Center for Transportation Research, University of
39 Texas at Austin, Austin, Texas, May 1993, 300 pp.
40 77. Nielson, M. P.; Braestrup, M. W.; Jensen, B. C.; and Bach, F., Concrete Plasticity, Beam
41 Shear – Shear in Joints – Punching Shear, Special Publication, Danish Society for
42 Structural Science and Engineering, Technical University of Denmark, Lyngby, 1978,
43 129 pp.

23
1 78. Ramirez, J.A., and Breen, J.E., “Evaluation of Modified Truss-Model Approach for
2 Beams in Shear,” ACI Structural Journal, Vol. 88, No. 5, September-October 1991,
3 pp. 562-571.

24
APPENDIX – DESIGN OF A DEEP BEAM REGION USING THE PROPOSED

STRUT-AND-TIE MODELING PROCEDURE

The following design example illustrates the use of the newly proposed strut-and-tie

modeling (STM) procedure for the design of a reinforced concrete transfer girder.

Description of Design Task

The transfer girder considered in this design example is shown in Fig. A1. The girder

has a total length of 29 ft (8839 mm) and supports a 20-by-20 in. (508-by-508 mm) column

(Column A). The girder, in turn, is supported by two 20-by-20 in. (508-by-508 mm) columns

assumed to behave essentially as simple supports (Columns B and C). The girder has a height,

h, of 5 ft (1524 mm) and a width, bw, of 2 ft (610 mm). As illustrated in Fig. A1, a factored

concentrated load, Pu, of 510.0 kips (2269 kN) acts on the girder from Column A. The

uniformly distributed loads on the member in addition to the member’s self-weight results in

a total factored load, wu, of 4.1 kip/ft (59.8 kN/m) acting along the length of the girder.

29’-0” [8839 mm]


20”
[508 mm]
[610 mm]

20”
2’-0”

[508 mm]

Pu = 510.0 kip [2269 kN]

20” x 20” Column


[508 mm x 508 mm] wu = 4.1 kip/ft [59.8 kN/m]
5’-0” [1524 mm]

CL Column A

20” x 20” Column 20” x 20” Column


[508 mm x 508 mm] [508 mm x 508 mm]

CL Column B CL Column C

1’-6” 6’-0” [1829 mm] 20’-0” [6096 mm] 1’-6”


[457 mm] [457 mm]

Fig. A1 – Transfer girder geometry and applied factored loads


Proposed Strut-and-Tie Modeling Method:
Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
25
The specified compressive strength of concrete, fc´, used for the design of the girder is

4000 psi (27.6 MPa). The specified yield strength of reinforcement, fy, is 60,000 psi (414

MPa).

Step 1: Separate B- and D-Regions

The first step in the STM design procedure is to separate the member into

discontinuity regions (D-regions) and Bernoulli regions (B-regions). D-regions are

characterized by a nonlinear distribution of strains through the member depth. In contrast, a

linear distribution of strains occurs within B-regions. According to St. Venant’s principle, an

elastic stress analysis indicates that a nonlinear stress distribution exists within about one

member depth from the location where a load or geometric discontinuity is introduced. In

other words, a linear distribution of stress can be assumed at about one member depth from

the discontinuity71. D-regions are therefore assumed to extend approximately a distance h

from the applied load and support reactions (i.e., discontinuities) in Fig. A2.

Column A
h = 5’-0” [1524 mm]

D-Region B-Region D-Region

Column B Column C

<h h h

Fig. A2 – Girder divided into D-regions and B-regions

A B-region exists within the right shear span of the girder (between Columns A and

C). The behavior of this portion of the member will therefore be dominated by the principles

of Bernoulli’s classic beam theory and can be designed using the sectional procedure. The

shear span at the left end of the girder (between Columns A and B) consists entirely of a D-

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
26
region and will be dominated by deep beam behavior. The sectional procedure is therefore

not appropriate for this portion of the member and the STM method should be used.

Alternatively, the shear span-to-depth ratio, a/d, for each of the shear spans of the girder can

be considered to determine the appropriate design procedures.

Step 2: Analyze Structural Component and Develop Strut-and-Tie Model

The load case presented in Fig. A1 should be modified to produce a loading for which

a strut-and-tie model can be developed. The loads applied to the STM must act only at the

nodes of the model; no external loads can be applied along the lengths of struts or ties. The

uniformly distributed load applied to the transfer girder is therefore divided into a set of point

loads that will act at the nodes of the STM. The set of loads shown acting on the member in

Fig. A3 are equivalent to the load case illustrated in Fig. A1. The 4.1 kip/ft (59.8 kN/m)

distributed load has been divided into five separate point loads. Two of these loads act along

the centerlines of Columns B and C. Another load acting at the location of Column A has

been added to the 510.0-kip (2269-kN) factored load, Pu, resulting in a total load of 536.0

kips (2384 kN). The strut-and-tie model will include three truss panels between the load at

Column A and the reaction at Column C. Therefore, two loads, each equal to 27.3 kips (121.6

kN), act within the right shear span.

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
27
1’-6” 1’-6”
[457 mm] 6’ -0” [1829 mm] 6’-8” [2032 mm] 6’-8” [2032 mm] 6’-8” [2032 mm] [457 mm]

C
L Column A

18.5 kip 536.0 kip 27.3 kip 27.3 kip 19.8 kip
[82.1 kN] [2384 kN] [121.6 kN] [121.6 kN] [88.1 kN]
5’-0” [1524 mm]

C
L Column B CL Column C

451.8 kip 177.1 kip


[2010 kN] [788.0 kN]

1’-6” 26’-0” [7925 mm] 1’-6”


[457 mm] [457 mm]

Fig. A3 – Point loads and support reactions that will be applied to the strut-and-tie model

Once all the point loads are applied to the girder, the simply-supported member is

analyzed to calculate the support reactions. The reactions at Columns B and C are determined

to be 451.8 kips (2010 kN) and 177.1 kips (788.0 kN), respectively.

The strut-and-tie model developed for the member is provided in Fig. A4. In this

figure, dashed lines represent struts, and solid lines represent ties. Although only the left

portion of the girder is designed using the STM procedure, a strut-and-tie model illustrating

the flow of forces from the loads to the reactions is developed for the entire member.
1’-6” 1’-6”
[457 mm] 6’-0” [1829 mm] 6’-8” [2032 mm] 6’-8” [2032 mm] 6’-8” [2032 mm] [457 mm]

C
L Column A

18.5 kip 536.0 kip 20” x 20” Column 27.3 kip 5.1” 27.3 kip 19.8 kip
[82.1 kN] [2384 kN] [508 mm x 508 mm] [121.6 kN] [130 mm] [121.6 kN] [88.1 kN]

A E
B -452.5 kip [-2013 kN] C -247.8 kip [-1102 kN] D
4’-2.8” [1290 mm]

[-82.1 kN]

[-88.1 kN]
-18.5 kip

-19.8 kip
[578.2 kN]
[456.6 kN]

130.0 kip
102.7 kip

F 35.21° 614.1 kip [2732 kN] 32.42° G 452.5 kip [2013 kN] H 247.8 kip [1102 kN]
I
32.42° 32.42°
451.8 kip 20” x 20” Column 4.1” 20” x 20” Column 177.1 kip
[2010 kN] [508 mm x 508 mm] [104 mm] [508 mm x 508 mm] [788.0 kN]

C
L Column B C
L Column C

Fig. A4 – Strut-and-tie model of the transfer girder

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
28
The first step in the development of the STM is to position the longitudinal tie located

along the bottom of the member (i.e., the bottom chord of the truss model). This tie represents

the force carried by the flexural tension reinforcement within the girder. The tie must

therefore be positioned to correspond with the centroid of the flexural reinforcement. Two

layers of No. 9 (Ø29 mm) bars are selected for the tie given the loads acting on the girder

(refer to Step 3 below). No. 4 (Ø13 mm) stirrups with 2 in. (51 mm) of clear cover will be

used, and a clear spacing of 1 in. (25 mm) will be provided between the two layers of No. 9

(Ø29 mm) bars. Therefore, the centroid of the longitudinal reinforcement is located at about

4.1 in. (104 mm) from the bottom surface of the member (refer to the final reinforcement

layout in Fig. A11).

Next, the placement of the horizontal strut along the top of the girder (i.e., the top

chord of the truss model) is determined. The strut is positioned to correspond with the

centroid of the equivalent rectangular compressive stress block as determined from a typical

flexural analysis. The depth of the rectangular stress block depends on the amount of flexural

tension reinforcement provided along the bottom of the member. Design iterations of the

STM procedure are likely necessary to determine the required flexural reinforcement and the

corresponding depth of the stress block. After a few iterations, 14-No. 9 (Ø29 mm) bars are

determined to be necessary to carry the longitudinal tie force (refer to Step 3 below). The

depth of the equivalent rectangular stress block, a, is therefore calculated as follows:

14 1.00 in.2 60,000 psi


10.29 in. 261 mm
0.85 0.85 4000 psi 24 in.

The horizontal strut is placed at a distance of a/2, or 5.1 in. (130 mm), from the top surface of

the girder.

The remaining struts and ties are positioned to model the flow of forces through the

depth of the member. Vertical struts are placed under the loads located directly above

Columns B and C to represent a direct load transfer to the supports. A diagonal strut is
Proposed Strut-and-Tie Modeling Method:
Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
29
extended between the applied column load at Node B and the support reaction at Node F to

model the deep beam behavior (i.e., bottle-shaped strut) in that region of the girder. Diagonal

struts and vertical ties are then added within the right shear span of the member. It should be

noted that Section A.2.5 of ACI 318-111 states that the angle between a strut and a tie

entering the same node shall not be taken as less than 25 degrees. A simple rule of thumb can

therefore be used to determine the appropriate number of truss panels to include between a

load and a support: the STM should include the least number of truss panels as possible while

still satisfying the 25-degree rule between the struts and ties entering a single node. Three

truss panels are provided between Nodes B and I to satisfy this condition.

Once the geometry of the STM is determined, the member forces of the struts and ties

are found by satisfying equilibrium. The resulting forces are shown in Fig. A4.

Recall that only the left shear span of the member will be designed using the strut-

and-tie modeling procedure. Thus, only the corresponding portion of the STM will therefore

be considered in the remainder of this example.

Step 3: Proportion Longitudinal Reinforcement

The next step in the design process is to proportion the longitudinal reinforcement to

satisfy the tie requirements of the STM. More specifically, the reinforcement required by Tie

FG must be determined. The strength of the reinforcement should be sufficient to carry the

calculated tie force. The area of steel, Ats, necessary to carry the 614.1-kip (2732-kN) factored

force in Tie FG is calculated as follows:

Factored load: 614.1 kips 2732 kN


Nominal tie strength:
Required area of reinforcement:
614.1 kips
, 13.65 in.2 8804 mm2
0.75 60 ksi

Using No. 9 (Ø29 mm) bars, each with a nominal area of 1.00 in.2 (645 mm2), 14 bars are

required and will be provided along the bottom of the girder. Recall from Step 2 that the

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
30
geometry of the STM is based on the amount of flexural reinforcement that is provided; thus,

the reinforcement determined in this step may result in slight modifications to the initially

assumed geometry.

Step 4: Determine Nodal Geometries and Perform Nodal Strength Checks

Within the deep beam region of the transfer girder, Nodes B and F must be checked to

ensure they have adequate strength to resist the imposed forces without the crushing of

concrete. Node B will be considered first.

A node with diagonal struts entering from both sides (i.e., from the right and from the

left) is generally divided into two parts in order to define the nodal geometry. Considering

Node B, Strut BF enters from the left side, and Strut BG enters from the right side. Thus, the

node should be divided. The applied 536.0-kip (2384-kN) load is split into two components

based on the percentage of the load that travels to each support, as illustrated in Fig. A5. The

length of the bearing face for each portion of Node B (refer to Fig. A6) is determined by

maintaining uniform pressure over the column area. These bearing face lengths are calculated

as follows:

751.6 kips sin 35.21°


20.0 in. 16.2 in. 411 mm
536.0 kips

191.5 kips sin 32.42°


20.0 in. 3.8 in. 97 mm
536.0 kips

The values used in the above calculations are labeled in Fig. A4 and/or Fig. A5. The line of

action for each component of the 536.0-kip (2384-kN) load is located at the center of each

respective bearing area [refer to Fig. A5(b)].

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
31
(a) (b)

Fig. A5 – Division of Node B: (a) original (global) STM; (b) STM with Node B divided into
two parts (1 kip = 4.448 kN)

The division of Node B into two parts causes a small change in the angle of Strut BF,

as shown in Fig. A5. The new angle of Strut BF is determined as follows:

50.8 in.
tan 35.93°
72.0 in. 20.0 in. 8.1 in.
2

where 50.8 in. (1290 mm) is the height of the strut-and-tie model measured between the top

and bottom chords of the truss, 72.0 in. (1829 mm) is the horizontal distance measured

between Nodes B and F, 20.0 in. (508 mm) is the dimension of Column A, and 8.1 in. (205

mm) is shown in Fig. A5(b). This small change in the strut angle minimally affects the

magnitude of the strut force acting at the node and is often neglected – adding conservatism

to the strength checks.

The geometries of both the left and right portions of Node B are illustrated in Fig. A6.

Recall that non-hydrostatic nodal geometries are used within the proposed STM procedure.

The length of the bearing face for each portion of Node B was previously calculated. The

height of the back face, hs, is taken as approximately equal to the depth of the equivalent

rectangular stress block [hs = 10.2 in. (259 mm)]. Calculation of the strut-to-node interface

length, ws, for the left portion of the node is provided in Fig. A6. The width of the node into
Proposed Strut-and-Tie Modeling Method:
Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
32
the page is taken as the dimension of Column A [20.0 in. (508 mm)] (refer to the strength

calculations below). The angle denoted “per global STM” in Fig. A6 is the angle of Strut BF

before the node was divided into two parts. Only compressive forces act on Node B, and it is

therefore classified as a CCC node (i.e., the concrete efficiency factors for CCC nodes are

used).

sin cos
16.2 in. sin35.93° 10.2 in. cos35.93° 
9.51 in. 8.26 in. 17.8 in. 451 mm

Fig. A6 – Node B

The 102.7-kip (456.6-kN) load applied to the right portion of the node is transferred

through the right shear span, which is designed using the sectional procedure. Therefore, the

stresses at the right portion of the node are not critical within the STM method, and only the

left portion of Node B will be considered. The geometry of the left portion of the node and

the factored forces acting on each face is clearly shown in Fig. A7. The force acting on the

back face is the sum of the horizontal components of Struts BC and BG and is calculated as

follows:

452.5 kips 191.5 kips cos32.42° 614.2 kips 2732 kN

where 32.42° is the angle of Strut BG in the global strut-and-tie model of Fig. A4.

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
33
Fig. A7 – Node B - left portion

Given that the girder is wider than Column A, triaxial confinement at Node B is taken

into account. The first step in evaluating the nodal strength is therefore to determine the

triaxial confinement modification factor, m, as illustrated in Fig. A8 and outlined in the

calculation below (refer to Section 10.14.1 of ACI 318-11):

24.0 in.
1.2 2 ∴ Use 1.2
20.0 in.

24.0” [610 mm] 20.0” x 20.0”


[508 mm x 508 mm]
B 2 Square Column, A1
45° 1

A2 is measured
on this plane
24.0” [610 mm]

20.0” x 20.0”
[508 mm x 508 mm]
Square Column,
A1 24.0” [610 mm]

Section B-B
45° through Beam
B
Top of Beam (Plan View)

Fig. A8 – Determination of the triaxial confinement modification factor, m, at Column A

The triaxial confinement factor, m, is applied to all faces of Node B according to the

proposed STM procedure. The bearing face of the left portion of the node is checked first.

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
34
The strength of the bearing face is calculated and compared to the applied load as shown

below (refer to Fig. A7):

BEARING FACE (NODE B – CCC)


Factored load: 433.3 kips 1927 kN
Concrete efficiency factor: 0.85
Effective compressive strength: 1.2 0.85 4.0 ksi 4.08 ksi
Factored nominal strength: 0.75 4.08 ksi 16.2 in. 20.0 in.
991.4 kips 433.3 kips
4410 kN 1927 kN

Therefore, the bearing face at Node B has adequate strength. Alternatively, the bearing

strength over the total 20-by-20 in. (508-by-508 mm) column area could have been compared

to the 510.0-kip (2269-kN) factored column load, Pu.

The strengths of the back face and strut-to-node interface of the left portion of Node B

are considered next. These two checks are presented below:

BACK FACE (NODE B – CCC)


Factored load: 614.2 kips 2732 kN
Concrete efficiency factor: 0.85
Effective compressive strength: 1.2 0.85 4.0 ksi 4.08 ksi
Factored nominal strength: 0.75 4.08 ksi 10.2 in. 20.0 in.
624.2 kips 614.2 kips
2777 kN 2732 kN

STRUT-TO-NODE INTERFACE (NODE B – CCC)


Factored load: 751.6 kips 3343 kN
Concrete efficiency factor: 0.85 4.0 ksi 20 ksi 0.65
Effective compressive strength: 1.2 0.65 4.0 ksi 3.12 ksi
Factored nominal strength: 0.75 3.12 ksi 17.8 in. 20.0 in.
833.0 kips 751.6 kips
3706 kN 3343 kN

The strength of Node B is therefore sufficient to resist the applied factored forces.

Node F is the other critical node within the deep beam portion of the girder. In order

to define the geometry of Node F, the number of forces acting on the node is reduced by

resolving Struts AF and BF into a single strut, as illustrated in Fig. A9. The force and angle

of the resulting strut are determined as follows:

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
35
18.5 kips 751.6 kips sin35.21° 751.6 kips cos35.21°
762.4 kips 3391 kN

18.5 kips 751.6 kips sin35.93°


tan 37.06°
751.6 kips cos35.93°

where 35.21° is the original angle of Strut BF per the global STM of Fig. A4. The original

angle of the STM is used to calculate the strut force in order to provide consistency with the

forces used in the strength checks of Node B. All other values in the above calculations are

shown in Fig. A9. Please note that 35.93° is the angle of Strut BF resulting from the division

of Node B.

Fig. A9 – Resolving Struts AF and BF (1 kip = 4.448 kN)

The geometry of Node F is illustrated in Fig. A10. The length of the bearing face, ls,

is equal to the full width of Column B, or 20.0 in. (508 mm). The height of the back face, ha,

is taken as twice the distance from the bottom surface of the girder to the centroid of the

longitudinal reinforcement (i.e., horizontal tie). Thus, the height of the back face is 2(4.1 in.)

= 8.2 in. (208 mm). The length of the strut-to-node interface, ws, is determined from the

calculation provided in Fig. A10. The width of the node into the page is again taken as the

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
36
column dimension. One tensile force acts on the node from Tie FG. Node F is therefore

classified as a CCT node.

sin cos
20.0 in. sin37.06° 8.2 in. cos37.06° 
12.05 in. 6.54 in. 18.6 in. 472 mm

Fig. A10 – Node F

The tie force at Node F results from the anchorage of the longitudinal reinforcing bars

(i.e., bonding stress) and does not concentrate at the back face. Thus, the back face does not

resist a direct force and is not critical provided the reinforcement is properly anchored (refer

to Step 6). A check of the back face is therefore unnecessary according to the proposed STM

procedure.

The strength checks for the bearing face and strut-to-node interface of Node F are

provided below. The triaxial confinement modification factor, m, is again equal to 1.2 (as

with Node B) and is applied to both faces of the node.

BEARING FACE (NODE F – CCT)


Factored load: 451.8 kips 2010 kN
Concrete efficiency factor: 0.70
Effective compressive strength: 1.2 0.70 4.0 ksi 3.36 ksi
Factored nominal strength: 0.75 3.36 ksi 20.0 in.
1008 kips 451.8 kips
4484 kN 2010 kN

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
37
STRUT-TO-NODE INTERFACE (NODE F – CCT)
Factored load: 762.4 kips 3391 kN
Concrete efficiency factor: 0.85 4.0 ksi 20 ksi 0.65
Effective compressive strength: 1.2 0.65 4.0 ksi 3.12 ksi
Factored nominal strength: 0.75 3.12 ksi 18.6 in. 20.0 in.
870.5 kips 762.4 kips
3872 kN 3391 kN

Thus, the strength of Node F is sufficient to resist the applied factored forces.

Step 5: Proportion Crack Control Reinforcement

Birrcher et al.67 recommend that 0.3% reinforcement be provided in each orthogonal

direction within the deep beam region of the girder to provide adequate strength and

serviceability performance. The reinforcement should be spaced evenly within the effective

strut area. Therefore, using two-legged No. 4 (Ø13 mm) stirrups, the required spacing of the

vertical crack control reinforcement is:

0.003 → 2 0.20 in.2 0.003 24 in.


5.56 in. 141 mm

Using No. 5 (Ø16 mm) bars as skin reinforcement, the required spacing of the horizontal

crack control reinforcement is:

0.003 → 2 0.31 in.2 0.003 24 in.


8.61 in. 219 mm

To satisfy these requirements, two legs of No. 4 (Ø13 mm) stirrups spaced at 5.0 in. (127

mm) will be provided, and No. 5 (Ø16 mm) bars spaced at 7.5 in. (191 mm) will be used as

skin reinforcement.

Step 6: Detail Reinforcement

Proper anchorage of Tie FG must be provided at the end of the girder (i.e., the

longitudinal bars must be developed at Node F). Also, reinforcement clear-spacing and clear-

cover requirements must be in accordance with the respective design standard. The final

reinforcement layout presented below satisfies the detailing provisions of ACI 318-11.

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
38
Reinforcement Layout

The final reinforcement details of the deep beam region of the transfer girder for the

load case considered in this design example are shown in Fig. A11 (please note that column

reinforcement is not shown).

Fig. A11 – Reinforcement details

Proposed Strut-and-Tie Modeling Method:


Tuchscherer, R. G.; Birrcher, D. B.; Williams, C. S.; Deschenes, D. J.; Bayrak, O. 4/2014
39
1 TABLES AND FIGURES

2 LIST OF TABLES:

3 Table 1 – Filtering of deep beam database.

4 Table 2 – Summary of stress checks used to evaluate deep beams

5 LIST OF FIGURES:

6 Fig. 1 – Characteristics of specimens in the evaluation database [N = 179]. (1 in. = 25.4 mm;
7 1 MPa = 145 psi)

8 Fig. 2 – Direct bottle-shaped strut bounded by a CCC and CCT node at each end.

9 Fig. 3 – Proportions for a) CCC and b) CCT nodal region.

10 Fig. 5 – Statistical comparison of the ratios of experimental to calculated capacity for


11 specimens in the evaluation database [N = 179].

12 Fig. 6 – Stress applied to the bearing [N=6] and back face [N=8] of a CCC node where said
13 surface controls capacity of specimen: Proposed Method. (1 MPa = 145 psi)

14 Fig. 7 – Stress applied to the bearing face of a CCT node [N=5] where said surface controls
15 capacity of specimen: Proposed Method. (1 MPa = 145 psi)

16 Fig. 8 – Stress applied to the CCC [N=53] and CCT [N=19] node interface where said
17 surface controls capacity of specimen. (1 MPa = 145 psi)

18 Fig. 9 – Stress applied to the strut-to-node interface where said surface controls the capacity
19 of the specimen: Proposed Method [N=160]. (1 MPa = 145 psi)

20 Fig. 10 – Influence that transverse reinforcement ratio has on the COV and conservatism of
21 various STM provisions.
22

40
1

2 Table 1 – Filtering of deep beam database.

No. Tests

Collection Database 905


Incomplete plate size info. -284
Stage 1 Subjected to uniform loading -7
Filtering Stub column failure -3
fc´ < 2000 psi -4
Filtered Database 606
bw < 4.5 in. -222
Stage 2 bw·d < 100 in2 -73
Filtering d < 12 in. -13
Σρ﬩ < 0.001 -120
Evaluation Database 179
3

4 Table 2 – Summary of stress checks used to evaluate deep beams.

Element Design Check Provision Allowable Stress


ACI 318 0.85f´c
Bearing AASHTO LRFD 0.85f´c
fib 0.85(1- f´c /40ksi)f´c
ACI 318 0.85f´c
CCC Node Back Face AASHTO LRFD 0.85f´c
fib 0.85(1- f´c /40ksi)f´c
ACI 318 0.64f´c
Node-Strut
AASHTO LRFD 0.85f´c
Interface
fib 0.85(1- f´c /40ksi)f´c
ACI 318 0.68f´c
Bearing AASHTO LRFD 0.75f´c
fib 0.70(1- f´c /40ksi)f´c
ACI 318 0.68f´c
CCT Node Back Face AASHTO LRFD 0.75f´c
fib N.A.
ACI 318 0.64f´c
Node-Strut
AASHTO LRFD f´c /(0.8+1701) ≤ 0.85f´c
Interface
fib 0.70(1- f´c /40ksi)f´c
Tie Tie All fy

41
1
120 120
Past studies (N = 144)

No. of specimens
100
No. of specimens 100
80 80
Birrcher et al. 2009 (N = 35)
60 60
40 40
20 20
0 0
14 22 30 38 46 54 62 > 70 4 8 12 16 20 24 28 32 36
Depth d, in. Width of web bw, in.
120 120
100

No. of specimens
100
No. of specimens

80 80
60 60
40 40
20 20
0 0
100 400 700 1000 1300 2 4 6 8 10 12 14 16 18
Area bw d, in2 Concrete compressive strength, ksi

120 120
100
No. of specimens

100
No. of specimens

80 80
60 60
40 40
20 20
0 0
0.0 1.0 2.0 3.0 4.0 0.1 0.25 0.4 0.55 0.7 >0.8

2 Longitudinal reinforcement l, %  v cos  + h sin ), %


3 Fig. 1 – Characteristics of specimens in the evaluation database [N = 179]. (1 in. = 25.4 mm;
4 1 MPa = 145 psi)
Bearing face
CCC node
Back
face
Node‐strut
interface
d

Node‐strut
interface
Back
face

CCT node
a ≤ 2d Bearing face
5
6 Fig. 2 – Direct bottle-shaped strut bounded by a CCC and CCT node at each end.

42
(1‐α) P αP

ll Line of action of member 
(1‐α) ll α ll forces
Bearing face Line separating width of 
node‐to‐strut interface
Back face
hs
α ll sin 
θ

Strut-to-node
a) CCC Node hs cos  interface

ha cos 
Strut-to-node
interface
θ ha
ls sin 

0.5ha
Back face
ls

b) CCT Node Bearing face


1
2 Fig. 3 – Proportions for a) CCC and b) CCT nodal region.
3
Hydrostatic nodes Non‐hydrostatic nodes

a/d = 1

a/d = 2
4
5 Fig. 4 – Influence of a/d on strut and node proportions.
6

43
ACI 318‐11 STM AASHTO LRFD (2010) STM
100 100
Min = 0.87 Min = 0.87
Conservative Max = 9.80 Conservative Max = 11.77
80 80
Mean = 1.81 Mean = 2.22
COV = 0.59
No. tests

60 60 COV = 0.69

40 40

20 20

0 0
0 1 2 3 4 5 6 7 8 9 10 11 12 0 1 2 3 4 5 6 7 8 9 10 11 12
Experimental/Calculated Experimental/Calculated
fib (1999) STM PROPOSED STM PROVISIONS
100 100
Min = 0.71 Min = 0.72
Conservative Max = 2.82 Conservative Max = 4.14
80 80
Mean = 1.51 Mean = 1.54
no. Tests

60 COV = 0.27 60 COV = 0.29

40 40

20 20

0 0
0 1 2 3 4 5 6 7 8 9 10 11 12 0 1 2 3 4 5 6 7 8 9 10 11 12
Experimental/Calculated Experimental/Calculated
1
2 Fig. 5 – Statistical comparison of the ratios of experimental to calculated capacity for
3 specimens in the evaluation database [N = 179].
2.0
CCC Back Face Controls
CCC Bearing Controls
1.5
0.85
Applied Stress

1.0
fc′

0.5

0.0
0 2000 4000 6000 8000 10000 12000
Concrete Compressive Strength (psi)
4
5 Fig. 6 – Stress applied to the bearing [N=6] and back face [N=8] of a CCC node where said
6 surface controls capacity of specimen: Proposed Method. (1 MPa = 145 psi)

44
2.0
CCT Bearing Controls

1.5
Applied Stress

0.70
1.0
fc′

0.5

0.0
0 2000 4000 6000 8000 10000 12000
Concrete Compressive Strength (psi)
1
2 Fig. 7 – Stress applied to the bearing face of a CCT node [N=5] where said surface controls
3 capacity of specimen: Proposed Method. (1 MPa = 145 psi)
ACI 318‐11
CCC Interface CCT Interface
[Min = 0.52, COV = 0.53]        [Min = 0.67, COV = 0.82]
2.0

1.6
Applied Stress

1.2
fc′

0.8

0.4

0.0
0 2000 4000 6000 8000 10000 12000 14000
4
5 Fig. 8 – Stress applied to the CCC [N=53] and CCT [N=19] node interface where said
6 surface controls capacity of specimen. (1 MPa = 145 psi)
7

45
2.0
CCC Strut-to-Node Interface
CCT Strut-to-Node Interface
1.5
Applied Stress

1.0
fc′

0.5
fc′
0.45 ≤  0.85 – ≤ 0.65
20 ksi
0.0
0 4000 8000 12000 16000
Concrete Compressive Strength (psi)
Normal Strength High Strength
1
2 Fig. 9 – Stress applied to the strut-to-node interface where said surface controls the capacity
3 of the specimen: Proposed Method [N=160]. (1 MPa = 145 psi)

ACI 318‐11 AASHTO LRFD (2010) fib (1999) PROPOSED


1
0.91
0.83
0.78
Coefficient of Variation (COV)

0.8
0.69 0.69 0.70

0.59
0.6 0.54

0.4 0.38
0.35
0.27 0.29 0.26
0.28 0.28 0.29

0.2

5 11 9 2 5 7 6 0 3 3 3 0 2 2 1 0
0
%
[N=179] [N=128] [N=45] [N=26]
# = number of unconservative estimations of strength
4
5 Fig. 10 – Influence that transverse reinforcement ratio has on the COV and conservatism of
6 various STM provisions.

46

View publication stats

You might also like