Sunteți pe pagina 1din 33

Special Moment Frames1

by Jack P. Moehle Pacific Earthquake Engineering Research Center University of California, Berkeley

1. Target Yield Mechanism


One of the guiding principles of seismic design is to spread yielding throughout the structure so that large inelastic deformations do not concentrate in isolated locations. In design of a Special Moment Resisting Frame (SMRF) it is important to avoid a yielding mechanism dominated by yielding of the columns in a single story, as this can result in very large local demands in the columns. Instead, it is desirable in a SMRF that yielding be predominantly in the beams (Figure 1.1). This is a fundamental objective in the design of a SMRF. Note that even if the beams are targeted as the main elements to yield, some column yielding must be anticipated. For example, yielding at the foundation seems likely (Figure 1.1). Also, it is difficult to completely protect the columns from yielding in other stories, as will be discussed later.

design lateral loads

Figure 1.1 Target yield mechanism

Given this capacity-design approach of having plastic hinges in the beams, the beams will be sized for the design seismic loads (usually based on analysis under code-specified loading), they will be detailed for ductile response, and the rest of the system will be proportioned to reduce the likelihood of inelastic action away from the beam plastic hinges.

2. Beams
2.1 Design objective
As discussed above, the design objective in a SMRF is to provide a stiff and strong spine of columns up the height of the building so that concentrations of inelastic action in isolated stories are avoided. Therefore, the objective is for beams to form flexural plastic hinges at targeted locations through the height of the frame. The design also should attempt to avoid inelastic response in shear as well as anchorage or bond failures.

1 Prepared for CE 244, Reinforced Concrete Structures, a graduate class taught at the University of California, Berkeley 1

2.2 Design actions on beams


The seismic demands imposed on building frames are a complicated function of the earthquake as well as the building stiffness, strength, mass, and configuration. Therefore, it is not possible to state with any accuracy the demands that beams in SMRFs, in general, need to sustain. However, some approximations of the level of inelasticity can be made. According to the International Building Code and Uniform Building Code, SMRFs are allowed to be designed for a force reduction factor of R = 8, that is, they are allowed to be designed for a base shear equal to one-eighth of the value obtained from elastic response analysis. Assuming an average building overstrength ratio of about 2.5 (actual building strength about 2.5 times the design value) owing to material overstrengths, section oversizing, strain-hardening, and interactions among structural components and among structural and nonstructural components that were not considered in design, the effective strength is about one-third of the strength required for elastic response. Accepting the equal displacement rule, the global displacement ductility for the building would be approximately 3. Local concentrations of interstory drift [Moehle, 1992] reasonably could result in local ductility demand about twice the global value, or equal to 6. For a typical beam span-to-depth ratio of ten, the local rotational ductility demand within the flexural plastic hinge would be approximately 2.5 times the local displacement ductility, resulting in an estimate of rotational ductility equal to 15. (See companion paper on seismic design principles for more detailed discussion of this topic.) An alternative approach views demands directly in terms of drift and member rotation demands. Assuming a global drift angle of 0.015 for a design-level event, the local drift ratio could reasonably be 0.03. The yield curvature for a typical beam is on the order of y/0.7h, and the flexural plastic hinge length can be approximated as being equal to h/2. Assuming that the gravity load results in moments at the beam ends less than half the moment capacity, and assuming that the columns do not contribute to drift (that is, they are relatively stiff), the curvature ductility for a given drift ratio can be calculated. For the drift ratio of 0.03, and for reasonable aspect ratios, the curvature ductility demand for a beam is approximately 20. This value is reasonably close to the value obtained in the preceding paragraph. Both approximations assume that the curvature demand is directly related to the building drift. This is only approximately correct, and even then only in the case of reversing plastic hinges. This subject is considered in more detail later.

2.3 Beam behavior


2.3.1 Formation of plastic hinges
An objective in design of SMRFs is to restrict most yielding to beams, which are specially detailed to resist the imposed actions. To get an understanding of the actions on a beam, consider the framing shown in Figure 2.1. Plastic moment strength can be assumed to be equal to the moment that will develop as the frame is deformed to the maximum deformation under the design earthquake loading. ACI 318 Chapter 21 defines Mpr as being equal to at least 1.25Mn. It is likely that larger moments will develop in a beam if it responds to the curvatures that are commonly anticipated for a SMRF. As discussed previously, curvature ductility demands on the order of 15 to 20 can be anticipated in some SMRFs. Under gravity loading shown in gray, the beam develops the moment diagram shown. Under service level gravity loads, it is expected that the moments will be less than half the nominal

moment strengths. Also, it is unlikely that the full service load will be present when the earthquake occurs, so even smaller moments are likely. If under earthquake loading the frame sways to the right, the column shears would be as shown by the black arrows. The moment diagram would shift under this loading, shown by the blue, dashed curve. The first section likely to reach the plastic moment is the negative-moment section. This section will begin to develop plastic rotation before the positive moment section yields. As the loading continues, the positive-moment section may yield and develop plastic rotation. While this is happening, the rotations at the negative-moment section continue to grow.

Mpr+ Mpry y

Bottom in tension

2.3.2 Behavior of reversing plastic hinges


As a building sways back and forth during an earthquake, the motion of the building drives the beam plastic hinges through displacement or deformation histories.

Envision the building drift history as shown by Figure 2.1 Plastic hinge formation the simple waveform (Figure 2.2). Assume the frame sways to point I. While this happens, the end of the beam that has negative moment (top in tension) under gravity loads is deformed as shown on the moment-curvature plot by the blue curve to point I. The strain distribution is also shown next to the beam, showing the top in tension, the bottom in compression note that the tensile strains exceed the compressive strains for a beam with modest amounts of tension longitudinal reinforcement. The stress-strain histories of the top and bottom bars also are shown below the beam. As the building drifts from point I to II (shown in red), the curvature of the beam reverses, as shown in the moment-curvature diagram. For the beam cross section shown, the cross-sectional area of the top reinforcement is larger than that of the bottom reinforcement. As a result, the total tension force developed by the beam bottom reinforcement is insufficient to yield the top reinforcement in compression. Therefore, unless other effects dominate (such as loss of bond through a beam-column joint and subsequent pull-through of the top reinforcement) the top beam reinforcement will be in compression but will not be strained to the yield point in this case the cracks that opened at state I remain open through the depth of the beam. If shear forces are high, this can lead to sliding along the cracked interface such sliding can lead to a moment-curvature relation that is pinched, with slackness occurring for low moment and curvature values. Note the stress-strain histories for the longitudinal reinforcement. The top reinforcement is in compression stress while showing tensile strain. Because the top reinforcement is in tension strain, the bottom reinforcement is driven to very large tensile strain in order to achieve the curvatures demanded of the beam. As the beam is flexed to position III, the situation reverses.

As
drift I II M I III III time

curvature II

III

A s
I fs III

II start here fs

II

s II III s

I
Figure 2.2 Reversing plastic hinge behavior

2.3.3 Reversing and non-reversing plastic hinges


If the beam is relatively short and/or the gravity loads relative low compared with seismic design effects, the beam behavior is likely to be as shown on the left of Figure 2.3. As the beam is deformed by the building response to the earthquake motions, the moments reach the plastic moment capacities at the beam ends (adjacent to the column face). As the earthquake sway reverses, the plastic hinges form again at the same locations, and a reversing plastic hinge forms, as presumed in the previous discussion. If the span or gravity loads are relatively large compared with earthquake effects, then a less desirable behavior can result. This is illustrated on the right-hand side of Figure 2.3. As the beam is deformed by the earthquake, the moments reach the plastic moment capacities in negative moment at the column face and in positive moment away from the column face. The deformed shape is shown. Upon reversal, the same situation occurs, but on opposite ends of the beam. In this case, the sections that had yielded previously do not yield in the opposite direction, but instead the plastic hinge forms in a different location. Note the deformed shape. As the deformations continue to reverse, the plastic hinges do not reverse, but instead continue to build up rotation. This results in progressively increasing rotations of the plastic hinges, so that for a long earthquake the rotations can be very large and the vertical movement of the floor can exceed serviceable values. This type of behavior should be avoided through design. In evaluation of existing buildings, this type of behavior should be investigated because if it does occur, the rotations can well exceed values estimated based on static inelastic analysis.

low gravity loads

high gravity loads

Mp+ Mp-

Mp + Mp -

reversing plastic hinge

non-reversing plastic hinge

Figure 2.3 Reversing and non-reversing plastic hinges

It is possible to determine whether non-reversing plastic hinges are likely. As shown in the two moment diagrams, reversing plastic hinges are expected if the slope of the positive moment diagram is negative, while non-reversing plastic hinges are anticipated if the slope is positive this of course wu assumes that the moment strength does Mpnot change appreciably along the span. Mp+ For the case of uniformly distributed load, consider equilibrium of the free VL body shown in Figure 2.4. Summing ln VR moments about VR:
+ M p + M p VL l n 2 wu l n =0 2

Figure 2.4 Free-body diagram cut through plastic hinges at end of beam

Setting the slope of the moment diagram (the shear VL) equal to zero, we find that a non-reversing plastic hinge is likely if
2 wu l n + Mp +Mp 2

2.3.4 Computed flexural ductility of beam cross sections


Response of beams in bending can be computed readily using computer programs (e.g., UCFyber). Most programs consider only monotonic loading response, though some are able to represent reversed cyclic loading effects. Regardless the program, the output is only as good as the input, the computation algorithm, and the ability of the user to make sensible interpretations.

To gain a sense of the important variables, Figure 2.5 presents results computed using UCFyber. Longitudinal reinforcement was assumed to have typical Grade 60 reinforcement properties, including strain-hardening. Concrete confinement was considered within the boundary of perimeter hoops, and followed the Mander relations. All other concrete was unconfined. Unconfined concrete had maximum strain capacity of 0.005. Confined concrete was assumed to have maximum compressive strain capacity of 0.02 (a practical limit for such sections considering reinforcement buckling and subsequent fracture). The sections were flexed so the top was in tension. Unconfined sections (Figure 2.5) were unable to reach the curvature ductilities estimated for SMRFs (approximately 15 to 20). Sections with the confinement were able to reach approximately the expected curvature ductility demand. ACI 318, Chapter 21, limits the ratio of area of tension reinforcement to area of compression reinforcement. Sections 1c and 1u represent cases that approach the ACI limit. For those sections, the available curvature ductility is barely equal to the required value even with transverse reinforcement.
No. 9 Grade 60 longitudinal bars, fc = 4 ksi
No. 3 stirrups @ 6 inches

18

24

Beam 1u
14000 12000 moment, kip-inch 10000 8000 6000 4000 2000 0 0

Beam 2u

Beam 1c

Beam 2c

y 20y

Beam 1c Beam 2c Beam 1u Beam 2u

0.001

0.002

0.003

0.004

0.005

curvature, 1/inch

Figure 2.5 Computed moment-curvature relations for unconfined and confined concrete sections. Top in tension.

2.3.5 Shear behavior


Consider a member subjected to a concentrated lateral load as shown in Figure 2.6. Assume that the 45-degree truss model effectively represents behavior in flexure and shear. Also, assume
6

that the tension chord is designed so that it does not yield, and that the compression diagonals are sufficiently strong that they do not crush. Yielding is controlled by yielding of the web reinforcement, shown in blue. In this case, on the first cycle of inelastic loading, the relation between reinforcement strain and applied shear will be nonlinear, as shown qualitatively in Figure 2.7. Upon unloading, the transverse reinforcement will unload and show a residual tension strain. Upon loading in the opposite direction, the transverse reinforcement will yield in tension again. Under repeated loading, the tension strain in the transverse reinforcement will tend to increase progressively. The result is lateral dilation of the member cross section, with widening cracks and eventual breakdown of the integrity of the concrete core. The ability to resist transverse loading decreases with continued loading. Reinforced concrete elements behave very differently depending on whether the inelastic action is predominantly in shear versus predominantly in flexure. The data shown in Figure 2.8 are from Jirsa [1977]. The upper loaddeformation relation is for a member where shear dominates; the lower relation is one where flexure dominates. Inelastic response in shear shows strength degradation associated with the shear-yielding mechanism of Figure 2.7. The flexural mechanism shows stable response for this configuration.

fy 0 web member tension stress

Figure 2.6 Truss model


transverse steel strain

Figure 2.7 Idealized strains in hoop reinforcement of shear-yielding member

2.4 Beam design


2.4.1 Longitudinal reinforcement
The ACI Building Code limits the longitudinal reinforcement ratio to 0.025. This limit is based on construction
Figure 2.8 Load-deformation response of shear-yielding and flexure-yielding members

considerations (it is difficult to place concrete with this much or more reinforcement), shear considerations (members with this much reinforcement tend to have excessively high shear stresses), and bond considerations (reinforcement must be anchored in joints, and this becomes increasingly difficult as the reinforcement ratio increases). Some codes limit the reinforcement ratio to some fraction of the balanced reinforcement ratio. Under reversed loading, where stresses are not uniquely related to strains (Figure 2.2), and in members with heavy hoop reinforcement (where concrete compression strains are enhanced), conditions are considerably different from those on which balanced failure computations are based, so this concept has been abandoned in the seismic provisions of the ACI code. As noted previously (Figure 2.2), if the cross-sectional areas of top and bottom longitudinal reinforcement differ significantly, cracks that open when the larger area of reinforcement yields will remain open on load reversal, unless the bars slip through the joint because of bond failure. To reduce consequences of this behavior, ACI limits the ratio of top to bottom reinforcement areas to between 0.5 and 2.0. The limits on reinforcement ratios also relate to conventional considerations of flexural ductility capacity. Figure 2.5 shows moment-curvature relations for some beam cross sections. When the ratio of tension to compression reinforcement areas differs significantly for large reinforcement ratios, the flexural ductility capacity is reduced. Confinement of the cross section with moderate transverse reinforcement considerably improves the computed behavior. Because of uncertainty in the moment requirements in seismic conditions as the frame responds to horizontal and vertical excitations, ACI 318 requires that the positive and negative moment strengths along the span be not less than one-fourth the strength provided at the face of the joint. Lap splices of longitudinal reinforcement are permitted in ACI 318 only if hoop reinforcement is provided over the lap length. The maximum spacing of the hoops is not to exceed d/4 or 4 inches. Laps are not to be used within joints, within a distance of twice the member depth of joints, and at locations where analysis indicates flexural yielding. The requirement for close spacing is based on the understanding that laps are effective only if closely spaced hoop reinforcement confines the splice after cover concrete spalls. Mechanical splices can be used, but should be Type 2.

2.4.2 Transverse reinforcement


Current building codes require special hoops to confine core concrete and restrain buckling of longitudinal reinforcement in plastic-hinge regions of beams. The special reinforcement is specified to extend a distance equal to twice the member depth from the center of the yielding region. The length of actual plastic hinge is likely to depend on the details of the moments and shears. One could argue that in the positive moment region the plastic hinge spreads over an extended length because the moment diagram is relatively flat (Figure 2.9). On the other hand, one could also argue that in the negative moment region the plastic hinge spread also is large because the plastic rotations are larger there and because the

low gravity loads

M p+ M p-

Figure 2.9 Moment diagrams of a yielding frame

shear forces and resulting tension shift are more significant there. The practical consideration is that it is difficult to identify the hinging zone with precision, and inexpensive to provide hoops over a generous length to cover the lack of precision. Hoop details within the target plastic-hinge region are designed to confine the core concrete so that it can reach strains well beyond the spalling strain and so that it can resist shear during these inelastic excursions; the close spacing of hoops also serves to restrain the longitudinal reinforcement from buckling when in compression. Specific details will depend on the configuration of the cross section. Figures 2.10 and 2.11 show some typical details of hoops as required by ACI 318-99. To restrain buckling, longitudinal reinforcement is to be restrained as required for columns in nonseismic construction, that is, corner bars are to be restrained in corners of hoops, at Detail A Detail B least alternate bars are to be restrained by crossties or intermediate hoop legs, and no unrestrained bar is to be more than 6 inches from a restrained bar. It is common in US practice to use crossties having 135-degree bends on one end and 90-degree bends on the other end. Although the 90-degree bends are known to be less effective in restraining buckling after loss of cover concrete (because the hook is not embedded in core concrete), it has been found to produce satisfactory behavior in members with low axial compression, and it improves contructibility. Cap ties sometimes are used (detail B in Figure 2.10) to ease construction. When used, the 90-degree hook should be restrained by an adjacent slab, as shown. Multiple hoops can be used as shown in detail C, but these details can be difficult to construct. New machinery for bending hoops is making it feasible to construct complex hoop configurations from a single piece of steel. These should be used where economical. Figure 2.12 is a photograph of beam reinforcement details in a building constructed in 2000 in Emeryville, California. The beam stood above the floor slab, so the details were visible for this photograph. According to the ACI 318 code, maximum spacing of hoop reinforcement within the plastic hinge region is the smallest of (a) d/4, (b) 8db, where db is the diameter of the

Detail C
Figure 2.10 Beam hoop details

Figure 2.11 Examples of overlapping hoops (/ACI 318)

longitudinal reinforcement, (c) 24 db, where db is the diameter of the transverse reinforcement, and (d) 12 inches. These requirements relate to the objective of confining the core concrete so it can resist shear under deformation reversals and can be strained to reasonable compression strains, as well as providing resistance to longitudinal bar buckling. Figure 2.5 shows computed behavior of cross sections with transverse reinforcement satisfying these provisions. Also, the beam is to be designed to resist the shear corresponding to development of Mpr at both ends of the member (Figure 2.13). Within the plastic-hinge region, when the shear due to seismic effects is equal to or greater than the gravity shear, the hoop reinforcement is to be designed to provide Vs assuming Vc = 0. Of course, this is not to imply that the concrete carries no shear, a misinterpretation that could mislead the engineer to think the concrete section is unimportant when in fact a stout section will improve section behavior. Instead, the intent is to increase the amount of hoop reinforcement to enable the concrete section to resist shear under adverse moment and shear reversals that are anticipated. Outside the plastic hinge region, stirrups are to be provided at spacing not to exceed d/2. Hoops are not required along this length. Figure 2.14 shows a beam detail with lap-spliced longitudinal reinforcement. The lap splices are placed near midspan, which is where the stresses are lowest for many SMRF frames (seismic actions predominate over gravity actions). Hoops are closely spaced near the ends (presuming flexural plastic hinges form at those locations). Hoops also are spaced closely along the lap splice to confine the splice.
Figure 2.12 Details in a building in Emeryville, CA

Figure 2.13 Design shears for girders and columns (ACI 318)

Rather than splice the reinforcement as shown in Figure 2.14, sometimes the reinforcement is curtailed at alternating positions along the span as shown in Figure 2.15. By cutting the bars in alternate spans, no lap splices are required.

10

Alternatively, mechanical splices can be used. Type 2 splices (capable of developing the specified bar ultimate tensile strength strength) can be placed anywhere along the span, though it is advisable to remove these from plastichinge regions.

seismic + gravity

gravity

Figure 2.14 Beam reinforcement with lap splices

Figure 2.15 Beam longitudinal reinforcement avoiding laps

3. Columns
3.1 Design objective
Design of a SMRF aims to achieve a beam-yield mechanism (Figure 1.1) and avoid a storyyield mechanism in which the columns in a story yield at the bottom and top of their clear length. Therefore, a capacity-design approach is used to promote flexural yielding in the beams and avoid flexural or shear yielding in columns. The capacity-design process begins by identifying where the inelastic action is intended to occur. For a SMRF, the inelastic action is intended to be predominantly in the form of flexural yielding of the beams. The building is analyzed under the design loads to determine the required flexural strengths of the beam plastic hinges. The beam sections are designed so that the reliable moment strength is at least equal to the design strength, that is, M n M u . Once the beam is proportioned, the plastic moment strengths of the beam can be determined based on the expected material properties and the selected cross section. ACI 318 uses the strength Mpr for this purpose. This moment is calculated using conventional ACI procedures with reinforcement yield stress taken equal to 1.25 times the nominal yield stress. The result is effectively the same as 1.25Mn. Knowing the beam plastic-moment strengths, it is possible to approach the problem of designing the columns so that they are stronger than the beams. The design problem is complicated by uncertainty in the seismic loading demands. Some aspects of the design problem are discussed in the following text.

11

3.2 Design actions on columns


3.2.1 Moments
Figure 3.1 shows schematically the beam and column moments that are obtained from a code analysis of a building frame under gravity and lateral loads. Beam moments vary nonlinearly because of the presence of distributed loads. Columns moments equilibrate the beam moments. Assuming no lateral inertial effect from the columns, the column moment diagrams are linear. The corresponding shears are constant along the height of a column in a story. The upper portion of Figure 3.2 shows a free-body diagram of a portion of a SMRF. The beams and columns have been cut at inflection points along the spans. The lower portion of Figure 3.2 shows the free-body diagram of the joint. Equilibrium of the joint requires the following relation:

column centerline beam centerline

Figure 3.1 Beam and column moments


Vc1

Vb1

(M c1 + M c 2 ) + (Vc1 + Vc 2 ) hb
2 = (M b1 + M b 2 ) + (Vb1 + Vb 2 ) hc 2
Mb1 Vc1 Vb1

Vc2 hc Mc1

Vb2

For simplicity in expression, this relation often is written as follows:

Mb2 Vb2 Vc2 Mc2

hb

= Mb

The latter expression is mathematically correct only if the beam and column dimensions are such that the terms involving the shears in the previous expression cancel. Codes commonly require that the sum of the column moment strengths exceed the sum of the beam moment strengths at joints. The intent of this requirement is to avoid formation of a story yield mechanism. Usually, the requirement is applied separately along each principal axis of the building. In three-dimensional SMRFs, loading along a diagonal can increase the moment demand on the columns. This is illustrated in Figure 3.3. Figure 3.3(a) shows a P-M interaction diagram for biaxial bending. A horizontal plane is located at the axial load corresponding to the loading considered. The plane cuts the P-M interaction surface as shown. This defines the column moment strength as a function of the direction of loading. For a circular column cross section with symmetric reinforcement,

Figure 3.2 Planar equilibrium


P design axial load corresponding Mx and My moment strengths Mx My (a) P-M interaction diagram Mb4

Mb1
f no y tio swa ec dir lding ui b

Mb3

Mb2

(b) Plan view of joint

Figure 3.3 Biaxial loading

12

the moment strength for bending along any direction is constant for a given axial load. For a square section with symmetric reinforcement, the moment strength for bending along the diagonal is somewhat less than for bending along any principal direction, the exact relation depending on the materials and reinforcement. Figure 3.3(b) shows an interior beam-column joint in plan. Moment vectors indicate the directions of the moments in the beams for building sway in the direction indicated. In this case, the sum of the moment strengths of the columns above and below the joint, for loading along the diagonal, must balance the vector sum of the beam moment strengths. For beams of equal strengths in the two directions, this can be expressed as

= Mb = 1

(M b1 + M b 2 + M b3 + M b 4 ) =

2 (M b1 + M b 3 )

Because the column moment strength is the same (for circular sections) or less (for square sections) for loading along the diagonal in comparison with loading along a principal axis, it is clear that the column demand is increased significantly by diagonal loading. The column moment strength must be increased by at least 40% for diagonal loading as compared with loading along a principal axis, if it is to be stronger than the adjacent beams. Dynamic response of SMRFs further complicates column moment design. As shown in Figure 3.4, the column moment diagrams follow a fairly regular pattern for first-mode loading (for a yielding system, modes do not exist in the sense defined by classical linear dynamics, but the concept of modes is convenient for discussion and design purposes). Note that even for this loading the moment diagrams are somewhat skewed in the bottom story because of the greater fixity provided by the foundation than by the first story above grade. For lateral force distributions that occur during dynamic response and that represent higher-mode loadings, the moment diagrams are less regular. Note that in some cases the column may not be in contraflexure. Note also that at some times (e.g., 2.73 seconds shown), the moments at a joint are carried almost entirely by either the column below or above the joint. Unless the column is much stronger than the beam, yielding in the column is likely.
first higher combined

(T. Kelly, U. Canterbury, 1974)

Figure 3.4 Higher-mode effects on column moment diagrams

moments

shear

3.2.2 Shears
Shear is equal to the slope of the moment diagram (Figure 3.5), so the moment patterns of Figure 3.4 indicate column shears as well as moments. In cases where
Case A Case B

Figure 3.5 Relation between shear and moment

13

the column moments are relatively large at both ends, a case that can be identified in some stories at time 2.73 in Figure 3.4, the shear also is relatively large. This case is shown as Case B in Figure 3.5. Therefore, the shears obtained from a first-mode or inverted triangular lateral loading are likely to underestimate the shears under dynamic response. The maximum column shear can be estimated by assuming formation of plastic moments at both ends, as shown in Figure 3.6; this shear may be unnecessarily large, so alternative estimates are often made. When using this approach, the axial load should be selected to obtain a conservatively high estimate of the column plastic-moment strength.

Pu Mpr Vu

Vu =

M pr

ln

ln

Vu Pu

Mpr

3.2.3 Axial Loads

Figure 3.6 Upper-bound column shear

Axial loads in SMRFs are the sum of the shears in the beams framing into the column plus the self weight of the column. As illustrated in Figure 3.7, the axial loads will vary around the building as a function of the column location and instantaneous earthquake loading. As shown, for earthquake and predominantly earthquake and earthquake inertial loading gravity effects gravity loads gravity effects from the left toward the right, opposite additive the gravity and earthquakeFigure 3.7 Gravity and earthquake axial loads on columns in induced axial loads will be different locations of a SMRF additive for the exterior columns on the right hand side of the building. The gravity and earthquake-induced axial loads will act in opposite directions for the exterior columns on the left-hand side of the building, and may result in column tension. For interior columns, the axial loads will tend to be dominated by gravity loads; variations due to earthquake loading will depend on the stiffness and strength of the beams framing into the column on either side this determines the shear forces in the beams and therefore determines the axial load variation. For regular buildings with equal spans and equal-size and strength beams in all spans, the axial load variation due to earthquake effects will be relatively small for interior columns. Limit analysis (plastic analysis) procedures can be used to calculate upper-bounds to column axial loads. For this purpose, a plastic mechanism is assumed for the building. For a SMRF, it may be reasonable to assume formation of beam flexural plastic hinges over the height of the building (Figure 3.8). With this assumption, the beam shears are summed over the height and added to the column weight to obtain the column axial force.

14

Corner columns tend to have the largest fluctuation in axial forces during an earthquake. This occurs for two reasons. First the corner columns usually support the smallest gravity loads. Second, for displacements along a diagonal direction, the beam shears are additive from both directions (Figure 3.9).

Vp Mpr Vp Vp Vp

Vp Mpr Vp Vp Vp

Vp Mpr Vp Vp Vp

Vp Mpr Vp Vp Vp

P = Vp + Wcolumn P = Vp + Wcolumn Plastic analysis gives an upperP = Vp + Wcolumn bound solution for column axial forces. Actual values may be less Figure 3.8 Plastic analysis to obtain upper-bound column depending on how much inelastic axial forces response develops in the frame and depending on how it develops. Building displaced Figures 3.8 and 3.9 assume that beam flexural plastic hinges toward corner form over the full height of the building. Studies show that column plastic hinging does not always form in this fashion. Figure 3.10 shows calculated results for a relatively flexible 12-story Vbeam building subjected to the 1940 El Center earthquake record. Vbeam As shown, beam flexural plastic hinges occur in discrete locations, those locations migrating up and down the frame as story drift concentrates in different stories at different times during the dynamic response.

The patterns shown in Figure 3.10 are highly dependent on the framing configuration, the dynamic characteristics of the frame, and the characteristics of the ground motion. For example, it is likely that a near-field ground motion pulse could impose massive sway in a frame that would result in nearly all the beams yielding at the same time. Even in this case, and even if the columns are made stronger than the beams, flexural yielding in general will not extend over the

Pcolumn

Figure 3.9 Corner column axial force

Figure 3.10 Calculated locations of beam flexural plastic hinges in a 12-story frame (T. Kelly, 1974)

15

full building height instead, flexural plastic hinges will form in the columns at some height that depends on the framing configuration, the relative column and beam strengths, and the distribution of lateral inertial loading.

3.3 Column behavior


Behavior of column cross sections in flexure is essentially the same as that of beams. Analysis generally follows the same assumptions and approach. The only significant difference for a column section is that the axial force is not necessarily equal to zero. Figure 3.11 shows a column cross section and associated analysis assumptions. Strain is assumed to vary linearly across the section, with maximum compression strain in concrete assumed to be the limiting parameter in some cases, fracture of longitudinal reinforcement is the critical behavior, but that will not be considered in detail here. Material stresses are assumed to be uniquely related to strains as discussed for beams, this assumption is not correct for inelastic reversed cyclic loading.
b d As d Atr As d As h c fs s fs Cs M P Cc Ts Ts

s s

fs

(a) Low axial load


s s s c fs fs fs P Cs M Cc Cs Ts

(b) High axial load


Figure 3.11 Flexural analysis of column cross sections

For relatively small axial force (Figure 3.11a), the depth of compression zone required to equilibrate the axial force is relatively small. Therefore, the curvature is relatively large when the maximum compressive strain capacity is reached. For larger axial force, the depth of compression zone required to equilibrate the axial force increases; therefore, the curvature capacity decreases. This effect of axial load on moment strength and curvature capacity is shown in Figure 3.12. The cross section shown is assumed to be unconfined. For unconfined sections, the flexural ductility decreases rapidly as axial load increases.

16

Figure 3.13 plots interaction diagrams for a column with 24 inch by 24 inch cross section reinforced with 16 No. 9 bars as longitudinal reinforcement. Concrete has compressive strength of 4000 psi, and reinforcement is Grade 60. The continuous curves are calculated behavior assuming the reinforcement yields at 60 ksi without strainhardening, and assuming concrete is unconfined.

strains

P
Higher Axial Balanced Axial Lower Axial
tension- compressioncontrolled controlled

Moment, M

Curvature,

Figure 3.12 Effect of axial load on flexural cross-section behavior

fc = 4 ksi #9 Grade 60 24 #3 @ 2.5

unconfined nominal confined nominal

Also shown in Figure confined strain hardening 3.13 are plots for two other assumptions. The broken curve is calculated 4000 response assuming 3000 reinforcement yields at 60 ksi without strain2000 hardening, and assuming 1000 confined concrete behavior. Transverse 0 reinforcement comprises a -1000 perimeter hoop plus cross tie of No 3 bars at -2000 longitudinal spacing of 2.5 0 5000 10000 15000 0 0.002 0.004 0.006 0.008 inches, which satisfies the Moment, kip-inch Curvature, 1/inch confinement requirements Figure 3.13 Calculated interaction diagrams of ACI 318. Confined concrete stress-strain relation is according to Manders model. Concrete confinement with transverse reinforcement increases the strain capacity and compressive stress capacity of the compression zone. This has the most significant impact on the P-M interaction diagram for axial forces exceeding the balanced point (the balanced point in fact changes and becomes ill-defined for confined concrete), as columns with high axial force are compression-controlled. However, the strength gain is marginal because the cover concrete has spalled, reducing the effective concrete section. Ultimate curvature capacity is increased for all axial loads. The dashed curve Figure 3.13 is for the same column cross section with confined concrete, but now assuming more realistic properties for the longitudinal reinforcement, including yield stress of 67 ksi, and strain-hardening to 110 ksi. This has the most significant effect on calculated behavior for low axial loads, as the column is tension-limited in this region. Note, however, that the effect of strain-hardening is realized mostly because the column has confined concrete, which increases the compression-zone strain capacity, thereby increasing the curvature capacity, resulting in increased tension reinforcement strain. Ultimate curvature capacity is reduced somewhat compared with the case of confined concrete with elasto-plastic reinforcement.
Axial Load, kip

17

3.4 Column design


3.4.1 Moment and Axial Load
According to US design codes, the reliable flexural strength of the column must be at least equal to the ultimate design moment. This expression is written as follows:

M n M u
Mu is the moment demand obtained from the code-required analysis of the building under the specified code earthquake representation. Furthermore, ACI 318 requires that the flexural strengths of the columns shall satisfy
Mc (6/5) Mg Mc = sum of moments at the faces of the joint corresponding to the nominal flexural strength of the columns framing into that joint. Column flexural strength is to be calculated for the factored axial force, consistent with the direction of the lateral forces considered, resulting in the lowest flexural strength. Mg = sum of moments at the faces of the joint corresponding to the nominal flexural strengths

of the girders framing into that joint. In T-beam construction, where the slab is in tension under moments at the face of the joint, slab reinforcement within an effective slab width defined in 8.10 is to be assumed to contribute to flexural strength if the slab reinforcement is developed at the critical section for flexure. Flexural strengths are summed such that the column moments oppose the beam moments. The column strength requirement needs to be checked for loading along the two principle directions of the frame, but it is permitted to consider one framing direction at a time. Note that this is not a joint equilibrium statement, but simply a requirement that the column moment strength exceed the girder moment strength by a set ratio. An alternative approach, which is more conservative, and which may be viewed as being more consistent with the capacitydesign philosophy, is to amplify the column design moments obtained from the code analysis on the basis of the expected flexural overstrength of the beams. The procedures is illustrated in Figure 3.14. Specifically, a flexural overstrength factor o is defined as
column centerline

Mpr,beam Mu,beam
beam outline beam centerline

Mu,column Mu,column

M = M

pr ,beam u ,beam

where Mpr,beam = the sum of probable moment strengths of the beams at the joint and Mu,beam = the sum of the code-required strengths of the beams at the joint (obtained from application of the code-specified design loading). The columns are then designed for

Figure 3.14 Column design moments

18

moments Mu,column = oMu,column, where Mu,column = the column moment obtained from the codespecified design loading. Regardless the method used to obtain moments, the axial loads should be checked for maximum and minimum values, as either may be critical, depending on the column configuration and the values of axial loads and moments. Figure 3.15 shows two loading cases, one with lateral load from the left, the other with lateral load from the right. The loadings are used to determine the axial loads Pmax and Pmin in the column shown. To obtain Pmax, the lateral load is shown from left to right, and a high estimate of the gravity load is imposed. The high estimate is obtained using a factored load combination that maximizes the axial load. In contrast, to obtain Pmin, the lateral load is applied from right to left, and a low estimate of the gravity load is imposed. In this latter case, use of a low estimate of gravity load results in a lower estimate of the axial load. As shown in the interaction diagram of Figure 3.15, both axial load cases need to be checked, as either one may be critical.
high gravity loads

increasing

Pmax
low gravity loads

Pmax Pmin M

Pmin
Figure 3.15 Axial load combinations for design

3.4 2 Transverse reinforcement


Transverse reinforcement serves to confine the concrete core, thereby increasing its compressive strain capacity for flexural and axial deformations and improving its toughness for resisting shear. Transverse reinforcement also helps to delay buckling of the longitudinal reinforcement and improves anchorage and splice strength. Aspects of shear strength are covered in the next section. The emphasis here is on transverse reinforcement for flexural and axial load enhancement, as well as buckling and anchorage resistance. Figure 3.16 idealizes the confining behavior of transverse reinforcement. As the core concrete is strained longitudinally, it expands laterally. The transverse reinforcement acts passively to restrain this dilation and thereby results in confining stress. To be most effective as confinement

19

reinforcement, transverse reinforcement should be evenly distributed along the length and around the perimeter. For circular-cross section columns, the circular hoops or spiral provides fairly uniform radial confinement because of the uniform curvature of the hoops or spirals. Close spacing of the circular hoops or spirals improves the uniformity of confinement along the length, and therefore improves confinement effectiveness. For rectangular-cross section columns, the rectangular hoops provide resistance effectively only at their corners or where crossties are placed. Therefore, confinement effectiveness is improved by placing longitudinal bars uniformly around the perimeter and providing hoops and crossties to restrain their outward movement. As with circular-cross section columns, confinement is improved by using relatively small longitudinal spacing. Figure 3.17 depicts confinement effectiveness relations obtained using the model proposed by Mander. The results are for a circularcross-section column, a squarecross-section column with two crossties in each direction, a square-cross-section column with one crosstie in each direction, and a square-crosssection column without crossties. The relation shown suggests that square-crosssection columns without crossties are much less effective than columns with crossties.
1.2 1.0

Circular Hoops
unconfined concrete effectively confined concrete

Rectangular Hoop

Figure 3.16 Confinement effectiveness

circular

Aeff/Acc

0.8 0.6 0.4 0.2 0.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
hoop + 1 crosstie each direction hoop without cross ties hoop + 2 crossties each direction

s/h

Figure 3.17 Confinement effectiveness relations for circular- and square-cross section columns

The specific requirements for transverse reinforcement for confinement depend on the specific demands imposed on the column. For example, a column framing into a foundation wall may be required to develop a plastic-hinge rotation equal to 0.01 or more for a design loading. Under this imposed rotation demand, the compression strain demand, and therefore the required amount of confinement reinforcement, will increase with increasing axial load. US codes for design of new buildings do not require computation of rotation demands. These demands are difficult to assess, especially in upper stories where the intent is to maintain an elastic column, but where yielding may occur owing to higher mode and multi-directional loading effects. Furthermore, axial loads occurring during earthquakes are difficult to determine during design. Prevailing design practice is to obtain the design axial load from the code-specified lateral forces; actual axial forces may deviate significantly from these values during an

20

earthquake. Plastic analysis procedures, which are not required by US codes, likewise may not produce an accurate estimate of axial loads. Given the uncertainty in rotation demands and axial force demands, ACI 318 specifies transverse reinforcement for confinement of critical regions that is independent of the level of axial force. The quantities required are intended to produce a column core that is capable of sustaining axial compression approximately equal to the axial compression capacity of the column before spalling of the concrete shell. This procedure has some advantages. It produces a column that is relatively tough and capable of sustaining axial forces due to unforeseen loadings that may crush the column; the result is a column highly resistant to brittle axial compression failure. The design, construction, and inspection processes are considerably simplified by using a constant amount of confinement up the height of the column. The required volume ratio of transverse reinforcement for circular cross sections is equal to

Ag f' 1 c 0.12 f c' f y s = 0.45 A f core y


Maximum spacing is not to exceed 3 inches. For rectangular cross sections, the total cross-sectional area of rectangular hoop reinforcement is not to be less than that required by either of the following two equations Ash = 0.3(shc fc/fyh)[(Ag /Ach) _ 1] Ash = 0.09shcfc/fyh Transverse reinforcement is to be provided by either single or overlapping hoops (Figure 3.18). Crossties are to be of the same bar size and spacing as the hoops, and each end of the crosstie must engage a peripheral longitudinal reinforcing bar. While it is recognized that 90-degree hooks on crossties are not as effective as 135-degree hooks, they are considered adequate for most loadings, and ease construction considerably. Some studies suggest that for heavy axial loads or
Atrfy
confining stress

Ash = 3Atr

hc

Atrfy

Atrfy

hc

Atrfy

Atrfy

confining stress

Atrfy

Atrfy

Ash = 4Atr

Figure 3.18 Rectangular hoops

21

unusually large column rotation demands, 135-degree hooks should be used on both ends. Where 90-degree hooks are used, consecutive crossties shall be alternated end for end along the longitudinal reinforcement. Perimeter hoops are required to have 135-degree hooks. Figure 3.18 shows how Ash and hc are defined for a rectangular-cross-section column. As suggested by Figure 3.17, the same confinement effectiveness can be achieved by different combinations of transverse and longitudinal spacings of hoop legs. ACI 318 recognizes this by specifying that the maximum longitudinal spacing sx vary as a function of the maximum horizontal spacing hx between legs of hoops or crossties around the column perimeter (Figure 3.19). In addition, longitudinal spacing of the transverse reinforcement is not to exceed (a) one-quarter of the minimum member dimension and (b) six times the diameter of the longitudinal reinforcement. The former requirements relates to confinement and to shear resistance requirements. The latter requirement is intended to restrain buckling of longitudinal reinforcement until deformations reach relatively large values.

sx 6 4

14

hx

Figure 3.19 Spacing limits for hoop reinforcement as function of horizontal spacing of hoop legs.

According to ACI 318, crossties or legs of overlapping hoops shall not be spaced more than 14 in. on center in the direction perpendicular to the longitudinal axis of the structural member. This longstanding limit is thought to be related to the intention to require crossties in significant columns, such columns perhaps being considered to have dimensions of 18 inches or larger. As suggested by Figure 3.17, crossties are desirable for confinement in all columns except very small and structurally inconsequential columns. Sometimes, architectural treatments result in thick concrete cover over the confined core. Such cases should be avoided where possible, as loss of unconfined cover can result in significant and relatively sudden loss of load-resisting capacity. As a minimum, ACI 318 requires that if the thickness of the concrete outside the confining transverse reinforcement exceeds 4 in., additional transverse reinforcement shall be provided at a spacing not exceeding 12 in. Concrete cover on the additional reinforcement shall not exceed 4 in. The transverse reinforcement described in the preceding paragraphs is to be provided over a length lo from each joint face and on both sides of any section where flexural yielding is likely to occur as a result of inelastic lateral displacements of the frame (Figure 3.20). The length lo is not to be less than (a) the depth of the member at the joint face or at the section where flexural yielding is likely to occur, (b) one-sixth of the clear span of the member, and (c) 18 in. Some studies suggest that the length of confinement at end regions should vary as a function of axial load level. Furthermore, as shown in Figure 3.4 and 3.20, moments in columns in the first story tend to be shifted toward the base, which may result in yielding extending over a greater length than in other typical stories. Some consideration should be given, therefore, to increasing the length lo in the lower stories. Also shown in Figure 3.20 is a typical detail at the first-floor level, where a slab on grade is situated on compacted soil or lean concrete. Usually the structural drawings for such details will include placement of a soft filler material between the column and slab on grade. If the fill is not placed, if it hardens or becomes contaminated with debris during construction or during the life of the building, or if the separation between the slab on grade and column is insufficient, the column

22

can be restrained by the slab on grade, shifting the inelastic action upward along the column height. This problem may have been a contributing factor in the critical damage of the Imperial County Services Building duringthe 1979 Imperial Valley California earthquake (Figure 3.21). Figure 3.20 suggests to measure the length from the top of the slab on grade, and extend the confinement down to the top of the footing.

lo
closely spaced hoops along lap splice

lo

lo

>lo

slab on grade

Where transverse reinforcement as specified above is not provided Moments throughout the full length of the column, the remainder of Figure 3.20 Moments and hoop spacing along column height the column length is to contain spiral or hoop reinforcement with center-to-center spacing not exceeding the smaller of six times the diameter of the longitudinal column bars or 6 in. The specification of relatively close spacing throughout the height is to avoid a sudden transition in toughness along the length, which could result in damage outside the heavily confined region. Mechanical splices of longitudinal reinforcement are permitted - Type 1 mechanical splices are not to be used within a distance equal to twice the member depth from the column or beam face or from sections where yielding of the reinforcement is likely to occur as a result of inelastic lateral displacements; Type 2 mechanical splices are permitted to be used at any location. Type 1 mechanical splices are those conforming to chapter 12 of ACI 318 (capable of 125% of the specified yield strength). Type 2 mechanical splices are capable of developing not less than the specified tensile strength of the spliced bar, so that they will be capable of developing significant inelastic strain without failure.

Figure 3.21 Imperial County Services building 1979 Imperial Valley earthquake

Lap splices of longitudinal reinforcement are permitted only within the center half of the member length, are to be designed as tension lap splices, and are to be enclosed within transverse reinforcement having longitudinal spacing not exceeding the minimum of (a) one-quarter of the minimum member dimension, (b) six times the diameter of the longitudinal reinforcement, and (c) sx defined by Figure 3.19. Horizontal spacing

23

between legs of hoops and cross ties is not to exceed 14 inches. In many cases, lap splices become prohibitively long, in which case mechanical splices are the only option. Figure 3.20 also suggests to avoid lap splices along the height of the first story. This suggestion is because the moment diagram is skewed so that the column may be subjected to significant moment at midheight where splices are usually located. Mechanical splices should be Type 2 in this story, though this is not required by the code. The details of the preceding paragraphs refer to columns of SMRFs. Columns supporting reactions from discontinued stiff members, such as walls, may be subjected to very large compressive forces that might result in column crushing. Such columns are to be provided with special confining reinforcement (the same as required within the length lo) over their full height beneath the level at which the discontinuity occurs if the factored axial compressive force in these members, related to earthquake effect, exceeds (Agfc/10). The transverse reinforcement is to exend into the discontinued member for at least the development length of the largest longitudinal reinforcement in the column. If the lower end of the column terminates on a wall, transverse reinforcement should extend into the wall for at least the development length of the largest longitudinal bar in the column at the point of termination. If the column terminates on a footing or mat, transverse reinforcement should extend at least 12 in. into the footing or mat.

3.4.3 Shear
According to ACI 318, the design shear Ve shall not be less than the factored shear determined by analysis of the structure under the code-specified loading Furthermore, the design shear force Ve shall not be less than the shear corresponding to development of a flexural plastic mechanism in the frame. One way of satisfying this requirement is to determine the design shear as the shear corresponding to development of the column probable moment strengths Mpr, where Mpr is the maximum probable moment strength calculated for the range of factored axial loads anticipated to be acting on the member. This condition is illustrated in Figure 3.22a, where a free body diagram of a column is shown in equilibrium with the column probable moment strengths acting at the column ends.
Pu Vu M Mpr V Mpr

Mpr

interaction curve calculated assuming steel has yield stress not less than 1.25 times the nominal yield stress.

ln

Pmax
Vu Pu Mpr Mpr V Mpr M
design moment and axial load combinations range of axial forces

(a) column hinging

(a) beam hinging

Pmin

Mpr

Figure 3.22 Column design shear

Figure 3.23 Determination of column Mpr

Figure 3.23 illustrates determination of the column probable moment strength Mpr. According to ACI 318, the P-M interaction curve is computed assuming the longitudinal reinforcement has stress capacity equal to not less than 1.25 times the nominal yield stress. The range of axial forces is determined from analysis of the building, either the code-specified lateral forces or a plastic mechanism analysis. The maximum plastic moment strength for that range of axial loads

24

is obtained not just for the maximum and minimum axial loads, but for all axial loads between those extremes. In some cases, the shear according to Figure 3.22a is significantly larger than the frame mechanism (intended to be a beam-yielding mechanism) can sustain, so the approach of the preceding paragraphs results in an excessively large design shear. Therefore, it is permitted to design for the shear determined from joint strengths based on the probable moment strength Mpr of the transverse members framing into the joint (Figure 3.22b). ACI 318 provides no guidance on how this shear is to be determined. As illustrated in Figure 3.5, if the moment is distributed unevenly between the columns above and below a joint, the shear in one column can be disproportionately larger than the shear in columns adjacent stories. Figure 3.24 illustrates one approach to satisfying the preceding requirement. The approach assumes that the frame mechanism is controlled by development of flexural plastic hinges in the beams. A flexural overstrength factor o is calculated as described previously. Correspondingly, the column moments are increased from Mu,column to oMu,column. Given that the shear is equal to the slope of the moment diagram, the shear is increased oVu,column, where Vu,column is the column shear calculated from the code-specified loading. The column shear should be determined from dynamic analysis rather than static analysis, so that higher-mode effects are approximately included. As noted previously, dynamic response results in shears higher than those that are obtained from a static lateral loading.

oVu,column Vu,column
Mu,column = o Mu,column Mu,column

Vu
Figure 3.24 Column shear design

The design shear strength will vary over the height of the column, in part because the transverse reinforcement varies, in part because the inelastic lo demands vary over the height. This is illustrated in Figure 3.25. Figure 3.25b shows shear strength over height for high axial loads. In this case, the strength is lo greater at the ends than along the middle because of the closer spacing of hoops along the length lo. Figure 3.25c shows the shear strength over height for low (a) column elevation (b) shear (c) shear strength strength for for low axial load axial load, or axial tension. Shear high axial load strength degrades with inelastic cyclic loading, and this degradation may be Figure 3.25 Shear strength especially pronounced for tension loading. According to ACI 318, transverse reinforcement along the length lo is to be proportioned to resist shear assuming Vc = 0 when both the following conditions occur: (a) The earthquake-induced shear force represents onehalf or more of the maximum required shear strength within those lengths; and (b) The factored axial compressive force including earthquake effects is less than Agfc/20.

25

4. Beam-Column Joints
The reader is referred to the ACI 352 report in addition to ACI 318.

4.1 Design objective


The design objective for reinforced concrete beam-column joints in SMRFs is for them to be stronger than the adjacent framing members so the majority of inelastic action occurs in the adjacent members, in particular, the beams. Sometimes this is referred to as keeping the joint elastic. However, considering yield penetration of beam reinforcement into the joint and slip of beam and column reinforcement from the joint, fully elastic behavior is not possible. Nonetheless, the design should aim to reduce these actions to acceptable levels through appropriate joint proportioning and reinforcement.

4.2 Design actions on joints


In a SMRF it is assumed that the design actions can be obtained from statics. The beam is assumed to develop flexural plastic hinges at the critical sections as shown in the right-hand side of Figure 4.1. The corresponding beam shear is obtained from statics. The beam moments and shears are applied to a free-body diagram of the column as shown in the lefthand side of Figure 4.1. For the purpose of determining the column shear for joint design it usually is sufficient to assume inflection points in the column at midheight, as shown. The column shears can subsequently be obtained by statics. The column shear force thus obtained may not be suitable for design of the column; however, it usually is sufficiently accurate for determining the design actions on the joint. For determination of the column shear for column design see previous discussion. Figure 4.2a shows a free body diagram of an interior joint. The beam moments have been replaced by the
Vc1

wu Mpr,b1 Mpr,b2 Vb2 Vb2 Vb1 Mpr,b1

Vb1

l nb

Vc2

Beam Section

Figure 4.1 Design actions on a joint

Tc1 Tb1 Vb1 Cb1 = Tb1

Vc1

Cc1 Tb1 Cb2 = Tb2 Vb2 Tb2

Vc1

Tb2

Vjoint
(b) Free body diagram to determine joint shear

Cc2

Vc2

Tc2

(a) Free body diagram showing external actions on joint

Figure 4.2 Free body diagrams of joint

26

corresponding internal force resultants (tension in longitudinal reinforcement and compression in concrete). It is assumed that the beam does not have axial force; therefore, the compression force resultant on one side of the joint is equal to the tension force resultant acting on the same side of the joint. For design purposes, ACI 318 assumes that the tension force is equal to 1.25Asfy, where As is the cross-sectional area of longitudinal reinforcement and fy is the nominal yield stress. The factor 1.25 is a low estimate of the overstrength for the reinforcement. According to ACI 318, As need not include the slab reinforcement in tension, though it would be conservative to do so. The column moment required to equilibrate the beam moments is obtained as the product of the column shear and half the column clear height (Figure 4.1). The internal forces associated with the combined moment and axial force (Figure 4.2a) need not be determined as part of the ACI design procedure; however, it is worth noting that the values are not necessarily equal because the column may have axial load. The design horizontal joint shear is obtained by equilibrium of horizontal forces acting on a free body diagram of the joint as shown in Figure 4.2b.

4.3 Joint behavior


Two simplified views of joint force-resisting mechanisms are shown in Figure 4.3 and 4.4. Figure 4.3 depicts what is referred to as a joint truss model. According to this model, longitudinal reinforcement tension and compression forces are resisted entirely by bond with the concrete. The bond produces diagonal compression forces in the joint concrete, which in turn are equilibrated by joint transverse reinforcement and longitudinal reinforcement.

1 3 2

For example, referring to Figure 4.3, the tension force in the beam longitudinal reinforcement entering the joint at the upper right hand corner of the joint results in bond forces along the bar length. At point 1, for Figure 4.3 Joint truss model example, this bond force (shown by green arrows) results in an inclined compression strut in the concrete (shown red). To equilibrate this at point 1 there also is required to be bond stress between the column longitudinal reinforcement and the concrete, as shown. At point 2, the diagonal compression strut is resisting by horizontal force in the joint horizontal reinforcement and by bond in the vertical column reinforcement. At point 3, the joint transverse reinforcement tension force is equilibrated by another diagonal compression strut in concrete along with bond stress in the column longitudinal reinforcement. At point 4 the diagonal compression strut is equilibrated by bond in the horizontal beam reinforcement as well as compression in the concrete. The truss model as described above requires high bond stresses between the longitudinal reinforcement and joint concrete, as it will be necessary to resist through bond both the longitudinal reinforcement tension force on one side of the joint plus the longitudinal reinforcement compression force from the other side of the joint. To resist this bond requires a large volume of joint horizontal reinforcement. Joint vertical reinforcement may be required to assist the column vertical bars in resisting the vertical forces in the joint. The ability of a joint to resist the bond stresses is hampered by a number of factors. At point 1, for example, the joint is

27

in tension due to flexural tension action of the adjacent column. Also, the length through the joint is not sufficient under normal conditions to develop the bar both in tension and compression. The problem is exacerbated by cyclic loading, which, among other things, deteriorates the bond capacity. An alternative model is the diagonal compression strut model, shown in Figure 4.4. According to this model, bond stress through the joint is lost under cyclic loading, so the reinforcement in tension on one side of the joint is developed through the compression zone on the opposite side of the joint, both within the joint and beyond the joint into the adjacent beam. This may result in the longitudinal reinforcement being in tension on both sides of the joint.

c b

For example, consider the beam flexural tension force at point a in Figure 4.4. Bond is likely to be substantial only along the compressed part of the joint near point b. The bond stress at this point is likely to be insufficient to fully develop the bar, so the bar Figure 4.4 Diagonal compression strut model remains in tension as it emerges from the joint at point c. Thus, the bar at point c is in tension even though the concrete in the flexural compression zone of the beam at that same point is in compression. Because both the top bars at c and bottom bars at d are in tension, the compression force in the concrete at c may be much higher than would be calculated from conventional flexural analysis. Confinement of the concrete in this region is important so that likelihood of brittle compressive failure is reduced. A result of the model shown in Figure 4.4 is that the joint shear is carried mostly through a diagonal compression strut as shown in yellow in the figure. Transverse reinforcement is required to confine the concrete. The transverse reinforcement strengthens the core concrete and toughens it so that it can undergo inelastic strain without failure. It also reduces dilation and crack opening under load reversals so that integrity of the concrete to resist compressive stresses is maintained. Note that both the truss model of Figure 4.3 and the diagonal compression stress model of Figure 4.4 require horizontal joint reinforcement, in one case to equilibrate shear forces and improve bond capacity, and in the other to enable the concrete to support the diagonal compression strut. The truss model would predict that the joint strength is directly related to the amount of transverse reinforcement, assuming that is the limiting component of the mechanism. The diagonal strut model does not relate the strength directly to the amount of transverse reinforcement. Also, note that both models require vertical reinforcement, in one case to resist vertical components of shear forces and to provide bond resistance, and in the other case to provide some vertical confinement. Either the column longitudinal reinforcement must be well distributed around the perimeter and remain elastic under lateral loading so it can act as the vertical joint reinforcement, or additional vertical steel must be added in the joint to resist dilation and vertical shear.

28

Some alternative models for joint shear strength (such as strut and tie models) can be used to avoid vertical or horizontal joint reinforcement, or both. These are not pursued here. Figure 4.5 shows test data on interior joints subjected to simulated seismic loading, as compiled by Otani [1991]. Open circles correspond to cases where the beam failed in flexure without apparent joint failure. Open squares correspond to cases where joints failed before beam yielding was apparent. Solid triangles correspond to cases beams first yielded, and then joints failed. The solid triangles follow a simple trend for low amounts of joint transverse reinforcement, the joint shear strength increases with increasing joint transverse reinforcement, but at a ratio of about 0.4 percent, the strength no longer increases for increasing joint transverse reinforcement ratio. This result has been interpreted as indicating that the truss model (Figure 4.3) does not model strength behavior very well, and has led to the use of the diagonal compression strut model (Figure 4.4) as the basis for design according to ACI 318. One of the outcomes of the ACI 318 design procedure (actually, the diagonal compression strut model) is that the beam longitudinal reinforcement may sustain bond deterioration that may result in inadequate performance as the reinforcement slips from (or slides through) the joint. This slippage can result in a behavior that is slack as the joint rotations pass near the origin. Figure 4.6 depicts load-deformation behaviors resulting from joints with different development lengths through the joint, identified by the ratio of the column dimension to the beam longitudinal bar diameter. Clearly the behavior is improved, with more stable and rounded hysteresis loops, as the ratio hc/ db increases. Note the slack or pinched behavior for small values of hc/db. Figure 4.5 shows that the strength of a joint can be well represented by the ratio of the horizontal joint shear stress to the compressive strength of the concrete. This is reasonable considering that failure is by failure of the diagonal compression strut (Figure 4.4). In US practices, it has been more traditional to relate the joint shear strength to the square root of the compressive strength of

Figure 4.5 Effect of transverse reinforcement quantity on joint shear strength

Figure 4.6 Effect of development length on behavior of beam-column joints.

29

concrete, based on the traditional notion that shear strength is a tension failure phenomenon that can be more directly related to the square root of the compressive strength rather than the directly to the compressive strength. Figure 4.7 shows test results for a joint designed with transverse reinforcement meeting the requirements of ACI 318, but with joint shear stress exceeding the allowable values of ACI 318 (to be described later) [Kurose, 1991]. The ACI 318 strength values are shown by Vu in the figures. The joint was subjected to loading in two horizontal directions, the upper loaddeformation relation for loading in one direction, the lower load-deformation relation for loading in the orthogonal direction. The joints were able to reach shear stress values exceeding the design value in both directions. However, under this overload the joints failed in shear. The deterioration in strength is apparent in the figures. Although the deformation capacity is considerable, by the 4% deformation cycles the strength deterioration probably is unacceptably high. Better behavior is obtained if the joint shear stresses are at or below acceptable limits.

Figure 4.7 Behavior of joints conforming with code details but having shear exceeding code values

4.4 Joint design


According to ACI 318, joint design requires (a) provision of adequate joint shear strength, (b) anchorage of reinforcement, (c) provision of adequate joint transverse reinforcement, and (d) provision of adequate column flexural strength.

4.4.1 Joint shear


The design horizontal joint shear is obtained by equilibrium of horizontal forces acting on a free body diagram of the joint as shown in Figure 4.2b. The design joint shear is not to exceed 0.85 times the nominal shear strength. The nominal shear strength of the joint is not tobe taken greater than the forces specified below for normalweight aggregate concrete. For lightweight aggregate concrete, the nominal shear strength of the joint is not to be taken greater than threequarters of the limits given for normalweight concrete. For joints confined on all four faces For joints confined on three faces or on two opposite faces For others 20 fc Aj 15 fc Aj 12 fc Aj

30

For the purpose of determining the appropriate joint shear strength, a member that frames into a face is considered to provide confinement to the joint if at least three-quarters of the face of the joint is covered by the framing member. Figure 4.8 summarizes the joint shear strength requirements. Note that joint geometries typical of construction at the roof level have not been included here. At the time of this writing, ACI Committee 352 is developing recommendations for joint shear strength for those geometries.

Vu Vn =
= 0.85
Type nonseismic seismic 24 20 20 15

f c' b j h

15 12

Values of

4.4.2 Anchorage of reinforcement


It is important for beam and column longitudinal reinforcement to be anchored adequately so that the joint can resist the beam and column moments.

Note: To qualify for a geometry shown, the beam must cover at least 75% of the face of the joint.

Figure 4.8 Summary of joint shear strength requirements


20d b

db

ldh

In interior joints, reinforcement typically extends through the Note: Provide at joint and is anchored in the least ldh, and always adjacent beam span (Figure 4.9). extend beam bars to near the back of the ACI 318 requires that the column joint dimension parallel to the beam longitudinal reinforcement be not (a) Requirement for (a) Requirement for less than 20 times the diameter of interior connections interior connections the largest longitudinal bar for normalweight concrete. For Figure 4.9 Anchorage/development requirements lightweight concrete, the required dimension is 26 times the bar diameter. This requirement helps improve performance of the joint by resisting slip of the beam bars through the joint. Some slip, however, will occur even with this column dimensional requirement. See Figure 4.6. ACI 352 recommends that the beam depth be not less than 20 times the diameter of the column longitudinal reinforcement for the same reason. ACI 318 does not include this requirement. For exterior joints, beam longitudinal reinforcement usually terminates in the joint with a standard hook. The provided length for a standard 90 deg hook in normalweight aggregate concrete must be the largest of 8db, 6 in., and the length required by the following expression:

l dh = f y d b / 65 fc

This expression assumes that the hook is embedded in a confined beam-column joint. The expression is limited to for bar sizes No. 3 through No. 11. For lightweight aggregate concrete,

31

the development length for a bar with a standard 90 deg hook is not to be less than the largest of 10db, 7-1/2 in., and 1.25 times that required by the above equation. In addition to providing the development length defined above, it is important to extend the hook to the far face of the confined column core. If the bars do not extend to the back of the joint, the full width of the joint will not participate in resisting the joint actions, as depicted in Figure 4.10. If a short development length is used, the joint ldh dimensions used for shear strength calculations should Back of be accordingly reduced. In heavily reinforced sections, joint not extending the tails of the hooks to the back face of the effectively confined core will create a wall of reinforcement that engaged may have negative effect on concrete integrity at that location. In those cases, the tails of the hooks can be withdrawn a short distance from the back of the cage.

4.4.3 Transverse reinforcement


Joint transverse reinforcement is provided to confine the joint core. As suggested by Figure 4.5, the strength Figure 4.10 Inappropriate anchorage is fairly independent of the amount of joint transverse of hooked bars reinforcement once a minimum amount is provided. According to ACI 318, the transverse hoop reinforcement in the joint is to be the same as the amount provided in the adjacent column end regions. Where members frame into all four sides of the joint and where each member width is at least three-fourths the column width, then within the depth of the shallowest framing member, transverse reinforcement may be reduced to one-half the amount required in the column end regions, so long as the maximum spacing does not Figure 4.11 Joint transverse reinforcement exceed 6 in.

4.4.5 Column strength


According to ACI 318, column moment strength is to exceed the beam moment strength. This requirement was discussed under the section on column design. This requirement serves two purposes. The first is for the column to be stronger than the beam so that a beam yielding mechanism will predominate, rather than a column yielding mechanism. Perhaps less critical, but still important, the longitudinal column reinforcement, being bonded to the joint concrete, serves as vertical joint reinforcement to restrain dilation. If the column longitudinal reinforcement yields, it is less effective for that purpose.

32

5. References
P.C. Cheung, T. Paulay, and R. Park, New Zealand Tests on Full-Scale Reinforced Concrete BeamColumn-Slab Subassemblages Designed for Earthquake Resistance, ACI SP 123, Design of BeamColumn Joints for Seismic Resistance, 1991. Y. Kurose, G.N. Guimaraes, L. Zuhua, M.E. Kreger and J.O. Jirsa Evaluation of Slab-Beam-Column Connections Subjected to Bidirectional Loading, ACI SP 123, Design of Beam-Column Joints for Seismic Resistance, 1991. Jirsa (Proceedings of a Workshop on Earthquake Resistant Reinforced Concrete Building Construction, UC Berkeley, 11-15 July 1977) Displacement-Based Design of RC Structures Subjected to Earthquakes, by J. P. Moehle, Earthquake Spectra, EERI, Vol. 8, No. 3, August 1992, pp 403-428. Otani, S. ACI SP 123, Design of Beam-Column Joints for Seismic Resistance, 1991.

33

S-ar putea să vă placă și