Sei sulla pagina 1di 9

ACI STRUCTURAL JOURNAL

Title no. 106-S45

TECHNICAL PAPER

Effective Stiffness of Reinforced Concrete Columns


by Kenneth J. Elwood and Marc O. Eberhard
Existing and proposed models of the effective stiffness of reinforced concrete columns subjected to lateral loads are assessed using the experimental response of 329 concrete columns. Existing models appropriate for design applications tend to overestimate the measured effective stiffness and are unacceptably inaccurate, because they generally neglect the influence of anchorage slip on the effective stiffness of the column. A three-component model that explicitly accounts for deformations due to flexure, shear, and anchorage-slip is shown to provide a more accurate estimate of the measured effective stiffness for the database columns. This model is simplified by neglecting small terms and approximating the results of moment-curvature analysis to obtain an accurate and rational effective stiffness model appropriate for design applications. For this model, the ratio of the measured stiffness to the calculated stiffness had a mean and coefficient of variation of 1.02 and 22% for circular columns and 0.95 and 25% for rectangular columns.
Keywords: bond slip; column; flexural stiffness; lateral loads; reinforced concrete; yield.

code, NZS 3101-06 (NZS 2006), based on the commonly used recommendations by Paulay and Priestley (1992). ACI 318-08 (ACI Committee 318 2008) will be the first edition of ACI 318 to provide stiffness recommendations specifically for lateral-load analysis. These code procedures are convenient for preliminary analysis, because they can be implemented without performing a moment-curvature analysis and without knowing the details of the column reinforcement. Simple effective stiffness models for application in design have also been proposed by Mehanny et al. (2001) and Khuntia and Ghosh (2004). This paper uses data from the Pacific Earthquake Engineering Research Center (PEER) Structural Performance Database, developed by the second author and his students (Berry et al. 2004), to assess the accuracy of these practical methodologies and to propose a new procedure. RESEARCH SIGNIFICANCE The assumed stiffness of a column dominates the results of linear and nonlinear analyses of buildings and bridges subjected to ground motions. Currently, most design professionals assume that the column effective stiffness is a fixed proportion (say 50 or 100%) of the gross-section stiffness. Using a database of 329 columns with rectangular and circular cross sections, this paper shows that existing procedures for estimating column stiffness are inaccurate. Based on simplifications of a three-component model, the paper proposes a new procedure that is more rational, practical, and accurate. The proposed procedure could be used immediately by design professionals, and it could be incorporated into design provisions, such as ASCE 41 (ASCE 2007a) or ACI 318 (ACI Committee 318 2008). DATABASE OF MEASURED EFFECTIVE STIFFNESSES The PEER Structural Performance Database (Berry et al. 2004) provided the data needed to evaluate the accuracy of various models of column stiffness. The database contains the cyclic force-deformation response, geometry, axial load, and material properties for more than 400 tests of reinforced concrete columns. A total of 366 of these columns were tested in cantilever, double-curvature, and double-cantilever configurations, which makes it possible to isolate the columns stiffness/flexibility from other sources of flexibility, such as a flexible supporting beam. To limit the analyses to columns typical of practice, the axial load was limited to a maximum of 0.66Ag fc, and the shear-span-to-depth ratio

INTRODUCTION The assumptions made in estimating the stiffnesses of structural members dominate the computed performance of a building or bridge subjected to earthquake ground motions. If these assumptions are used in a linear analysis, they control predictions of the period of the structure, the distribution of loads within the structure, and the deformation demands. The member stiffnesses also control the yield displacement, which in turn affects the displacement ductility demands calculated as part of a nonlinear analysis. The consequences of overestimating or underestimating the actual stiffnesses of structural members depend on the type of structural system and the response parameter of interest. For example, a low estimate of the effective stiffnesses of columns in a moment-resisting frame usually leads to a conservative (high) estimate of the displacement demands. In contrast, a low estimate of the effective stiffnesses for columns in a shear-wall building would lead the designer to unconservatively underestimate the elastic shear demands on the columns. The need for an accurate estimate of effective stiffnesses is even more crucial for time-history analyses, in which the peaks and valleys of the ground-motion response spectrum significantly influence the computed performance. To assist engineers in developing numerical models for the estimation of lateral deformation demands, most codes and standards provide recommendations for member effective stiffness. The Federal Emergency Management Agency (FEMA) 356 seismic rehabilitation guidelines (ASCE 2000) specify the most commonly used procedure for estimating column stiffness in the U.S. This procedure has been adopted into the Seismic Rehabilitation Standard, ASCE 41 (ASCE 2007a). It recently has been superseded by a new procedure specified in ASCE 41 Supplement 1 (ASCE 2007b), described in Elwood et al. (2007). A similar model is included in the commentary to the New Zealand concrete 476

ACI Structural Journal, V. 106, No. 4, July-August 2009. MS No. S-2007-399.R1 received July 8, 2008, and reviewed under Institute publication policies. Copyright 2009, American Concrete Institute. All rights reserved, including the making of copies unless permission is obtained from the copyright proprietors. Pertinent discussion including authors closure, if any, will be published in the May-June 2010 ACI Structural Journal if the discussion is received by January 1, 2010.

ACI Structural Journal/July-August 2009

ACI member Kenneth J. Elwood is an Associate Professor at the University of British Columbia, Vancouver, BC, Canada. He received his PhD from the University of California, Berkeley, Berkeley, CA. His research interests include the behavior and performance-based design of reinforced concrete structures under seismic loading. He is Chair of ACI Committee 369, Seismic Repair and Rehabilitation, and a member of ACI Committee 374, Performance-Based Seismic Design of Concrete Buildings, and Joint ACI-ASCE Committee 441, Reinforced Concrete Columns. Marc O. Eberhard, FACI, is a Professor at the University of Washington, Seattle, WA. He received his BS in civil engineering, and materials science and engineering from the University of California, Berkeley, and his MSCE and PhD from the University of Illinois, Chicago, IL. He is a member of Joint ACI-ASCE Committee 445, Shear and Torsion.

was limited to a minimum of 1.4 (all variables are defined in the Notation section). The selected specimens include 221 columns with rectangular cross sections with rectangular transverse reinforcement and 108 columns with circular, octagonal, and square cross sections with spiral transverse reinforcement (for brevity, these database subsets will be referred to as rectangular columns and circular columns, respectively). The minimum, maximum, mean, and median properties of the selected column databases are reported in Table 1. Both sets of column data have wide ranges of column parameters; and for both sets of data, the strength of over 90% of the included columns was limited by their flexural strength. This statistic is a consequence of the limit on the aspect ratio, which eliminated many shear-critical columns. The distribution of some parameters varies greatly between the two sets. For example, compared with the rectangular column dataset, the circular column set includes fewer columns with axial loads above 0.3Ag fc or fc above 60 MPa (8700 psi). The circular columns also tend to have lower values for db /D when compared with the rectangular columns. These differences

are consistent with the observation that many of the tests of circular columns were proportioned to represent typical bridge practice, whereas most of the rectangular columns were proportioned to reflect construction practice for buildings. For each column, the envelope of the measured lateral load-displacement relationship was corrected for P-delta effects to give the effective lateral force envelope for each column. The yield displacement and effective stiffness of each column were determined as shown in Fig. 1. For approximately 90% of the columns, the effective stiffness was defined based on the point on the measured effective force-displacement envelope that corresponded to the calculated force at first yield, Ffirst yield (Fig. 1(a)). Adopting the same definition of yield used by Benzoni et al. (1996) and others, the yield force was defined as the first point at which the tension reinforcement yielded or the maximum concrete strain reached a value of 0.002, whichever came first. The database does not include strain measurements, so the force at first yield was determined with a moment-curvature analysis, using a linear model for the steel and the Mander et al. (1988) constitutive relationship for the concrete. This definition could not be used for columns whose strength did not substantially exceed the yield force (for example, shear failures). For these columns, defined as those whose maximum measured effective force Fmax was not at least 7% larger than the calculated force at first yield, the effective stiffness was defined based on the point on the measured force-displacement envelope with an effective force equal to 0.8Fmax (Fig. 1(b)). Assuming the column is fixed against rotation at both ends and has a linear variation in curvature over the height of the column,

Fig. 1Definition of yield displacement and effective stiffness from test data for: (a) yielding columns; and (b) columns that did not yield. Table 1Range of properties for database columns
Circular columns Rectangular columns (108 specimens) (221 specimens) Mini- MaxiMini- MaxiParameter mum mum Mean Median mum mum Mean Median a/D fc, MPa (ksi) fy, MPa (ksi) D/db v f c , MPa (psi) P/Agfc 1.5 18.9 (2.7) 10.0 4.0 4.0 34.4 (5.0) 1.5 21.0 (3.0) 7.6 3.6 3.2 36.5 (5.3) 90.0 37.9 (13.1) (5.5) 118.0 52.3 (17.1) (7.6)

240 565 420 446 318 587 456 453 (34.8) (82.0) (61.0) (64.7) (46.2) (85.2) (66.2) (65.7) 0.005 12 0.05 (0.6) 0.1 0.056 0.024 0.021 48 27 28 0.19 (2.3) 0.10 0.010 12 0.09 (1.1) 0.00 0.060 0.024 0.021 32 0.71 (8.6) 0.63 18 0.32 (3.9) 0.23 16 0.30 (3.6) 0.20

0.99 0.27 (11.9) (3.3) 0.58 0.15

Fig. 2Effect of key parameters on measured effective stiffness. 477

ACI Structural Journal/July-August 2009

the measured effective modulus of rigidity (for simplicity, referred to here as stiffness) can be defined as F 0.004 a EI effmeas = -----------------3 y
3

(1)

where F0.004 is the calculated effective force on the column when the extreme concrete fiber reaches a maximum compressive strain of 0.004, and y is the displacement at yield according to Fig. 1 for an equivalent cantilever of length a. Figure 2 demonstrates the influence of several key parameters on the measured effective stiffness, expressed as a fraction of the gross-section stiffness, EIg. Within the dataset, the

measured effective stiffness ranges from 10 to 122% of the gross-section stiffness (Table 2). The measured effective stiffness can exceed EIg because the effect of the longitudinal reinforcement on the transformed section is not accounted for when determining EIg. Correlation coefficients for the rectangular and circular columns (Rr and Rc , respectively) are also shown in Fig. 2. The normalized effective stiffness increases most consistently with increasing axial-load ratio (P/Ag fc) and aspect ratio (a/D), with correlation coefficients ranging from 0.58 to 0.79. The normalized effective stiffness also increases with increases in concrete compressive strength (fc). The normalized stiffness decreases with an increase in the ratio of the steel yield stress to concrete compressive strength (fy /fc). The normalized stiffness correlates only weakly with the normalized bar size and longitudinal reinforcement ratio. EVALUATION OF EXISTING MODELS The models implemented in many of the structural codes are similar in form to each other. Chapter 8 of ACI 318-08 (ACI Committee 318 2008) provides three options for approximating member stiffnesses for the determination of lateral deflection of building systems subjected to factored lateral loads: (a) 0.35EIg for flexural members (P < 0.1Ag fc) and 0.7EIg for compression members (P 0.1Ag fc ); (b) 0.5EIg for all members; or (c) as determined by a more detailed analysis considering the reduced stiffness of all members under the loading conditions. Figure 3 and Table 2 compare options (a) and (b) with the measured effective stiffnesses from the column databases. Option (a) generally overestimates the effective stiffness for axial loads below 0.4Ag fc. Option (b) overestimates the stiffness for columns with low axial loads and underestimates the stiffness for columns with high axial loads. The figure and table also include evaluations of the effective stiffness models from FEMA 356 (ASCE 2000), ASCE 41 Supplement 1 (ASCE 2007b), and Paulay and Priestley (1992), all of which allow for interpolation between effective stiffness values at low and high axial loads. The FEMA 356 (ASCE 2000), and the Paulay and Priestley (1992) recommendations also tend to overestimate the measured effective stiffness for the columns with low axial loads, particularly for the rectangular column dataset. Of these existing procedures, ASCE 41 Supplement 1 (ASCE 2007b) provides the best average estimate of the measured effective stiffness (refer to Table 2). But none of these models are accurate. The coefficient of variation for all of these models ranges from 35 to 58%, depending on the particular model and dataset. As will be demonstrated in the following, this scatter can be

Fig. 3Measured effective stiffness from database compared with existing code models (ASCE 41-S1 = ASCE Supplement 1 [ASCE 2007b]; PP92 = Paulay and Priestley [1992]).

Table 2Statistics for ratio of measured to calculated effective stiffness for existing effective stiffness models
Circular columns (108 specimens) Model Gross section FEMA 356 ASCE 41 Supplement 1 ACI 318-08 (a) ACI 318-08 (b) Paulay and Priestley (1992) Mehanny et al. (2001) Khuntai and Ghosh (2004) Minimum Maximum 0.13 0.25 0.42 0.24 0.25 0.26 0.26 0.21 1.21 1.96 2.11 1.81 2.42 1.62 1.24 1.58 Mean 0.39 0.74 1.02 0.76 0.78 0.68 0.61 0.68 Median 0.34 0.64 0.91 0.69 0.67 0.61 0.56 0.62 Coefficient of variation, % 55.1 48.1 39.1 46.7 55.1 41.3 37.7 41.8 Minimum 0.10 0.19 0.27 0.14 0.19 0.17 0.16 0.16 Rectangular columns (221 specimens) Maximum 1.22 1.95 1.95 1.74 2.43 1.56 1.29 2.03 Mean 0.37 0.68 0.82 0.59 0.75 0.58 0.49 0.66 Median 0.33 0.64 0.83 0.53 0.65 0.56 0.49 0.62 Coefficient of variation, % 58.1 48.8 36.0 49.5 58.1 44.4 41.2 54.3

478

ACI Structural Journal/July-August 2009

attributed mainly to the fact that these procedures are based on consideration of expected flexural deformations. Using a computed moment-curvature relationship, as shown in Fig. 4, the effective flexural stiffness of the column EIflex can be determined based on the moment at first yield of the column, Mfirst yield. The moment-curvature response was determined for each column in the database based on plane-section analysis and using the concrete constitutive model by Mander et al. (1988) and a linear constitutive model for steel. Figure 5 compares the code-based models with the calculated flexural stiffnesses of the columns in the database expressed as a fraction of the gross-section stiffness (EIflex /EIg). For many of the columns considered, the models provide an adequate estimate of the flexural stiffness. Comparing the results in Fig. 3 and 5, however, it is apparent that other sources of flexibility must be taken into account to accurately estimate the total effective stiffness. Other effective stiffness models incorporating the influence of variables beyond axial load have been proposed in the literature. For example, Mehanny et al. (2001) accounts for the influence of the longitudinal reinforcement by introducing a model based on the transformed moment of inertia and the balanced axial load EI effcalc EI g, tr = ( 0.4 + P 2.4P b ) 0.9 (2)

THREE-COMPONENT MODEL OF YIELD DISPLACEMENT Several researchers (Sozen 1974; Priestley et al. 1996; Lehman and Moehle 1998; Berry and Eberhard 2007) have proposed estimating the yield displacement of an equivalent cantilever column of length a as the sum of the displacement components due to flexure, shear, and bar slip y = flex + shear + slip (4)

Fig. 4Definition of yield curvature and flexural stiffness (modulus of rigidity).

The statistics reported in Table 2 indicate that the Mehanny model consistently overestimates the measured effective stiffnesses and has large coefficients of variation (38% for circular columns and 41% for rectangular columns). Khuntai and Ghosh (2004) recommend an effective stiffness model for lateral-load analysis of reinforced concrete frames, with and without slender columns, accounting for the influence of longitudinal reinforcement and effective eccentricity of the axial load. They propose the following equation for compression members (P > 0.1Ag fc )
EI effcalc EI g = ( 0.80 + 25 ) ( 1 e D 0.5P P o ) 1.0

(3)

The effective stiffness from Eq. (3) is limited to greater than the effective stiffness for flexural members determined based on a similar model included in Khuntai and Ghosh (2004). Compared with the Mehanny model, the Khuntai and Ghosh model better predicts the average stiffness, but the coefficient of variation for the ratio of the measured to calculated effective stiffnesses exceeds 40% for the circular columns and exceeds 50% for the rectangular column dataset (Table 2). With the exception of ASCE 41 Supplement 1 (ASCE 2007b), all of the models considered were developed primarily to provide an estimate to the flexural effective stiffness (determined based on moment-curvature analyses) and, hence, ignore additional flexibility due to bar slip and shear deformations. Consequently, these models ignore the important dependence of the effective stiffness on the aspect ratio of the column, evident in Fig. 2(b). Rather than relying on purely statistical models, it would be preferable to develop a simple model, whose form is based on the theoretical calculation of the yield displacement accounting for the flexibility due to flexure, shear, and bar slip. As shown in the following sections, this approach can provide a more accurate estimate of the measured effective stiffnesses of the database columns. ACI Structural Journal/July-August 2009

Fig. 5Comparison of flexural stiffness with code models (ASCE 41-S1 = ASCE Supplement 1 [ASCE 2007b]; PP92 = Paulay and Priestley [1992]). 479

In this section, a similar model is developed based on the column datasets. Then, the three components of deformation are combined into a single, nondimensional equation, which can serve as the basis for the development of practical effective-stiffness models based on gross-section properties. Flexural deformations The calculated flexural curvatures in a reinforced concrete column can be integrated directly to estimate the column deformations attributable to flexure. Alternately, assuming a linear variation in curvature over the height of the column, the contribution of flexural deformations to the displacement at yield can be estimated as follows aa - M 0.004 flex = ---- y = ---- ------------3 3 EI flex
2 2

Bar slip deformations Slip of the reinforcing bars within the beam-column joints or foundations further increases the lateral displacements. This section derives an expression to estimate the lateral displacement of a column due to bar slip prior to yielding of the longitudinal reinforcement. Moments at the ends of a reinforced concrete column tend to cause tension in the longitudinal reinforcing bars, as shown in Fig. 6. This tension force Ts must be resisted by the bond stress u between the reinforcement and the footing or joint concrete. If the bond stress is assumed to be constant, equilibrium considerations lead to the following expression for the length of bar required to resist Ts db f s l = -------4u (7)

(5)

where M0.004 is the flexural moment at a maximum concrete compressive strain of 0.004, and y is the yield curvature, as defined in Fig. 4. Shear deformations The column deformation due to shear within the elastic range of response is small for most columns, but it can be large (relative to others sources of deformation) for stocky columns with high levels of shear demand. Before shear cracking, this contribution can be estimated by assuming that the effective shear modulus is equal to the gross-section, isotropic elastic value (G = Ec /2.4). As the shear cracking increases, the effective shear modulus reduces significantly. For many applications, it is convenient to estimate the shear displacement of an equivalent cantilever column by idealizing the column as a homogeneous, isotropic material with a constant, reduced shear modulus M 0.004 shear = -------------A v G eff (6)

Using Eq. (7) and integrating the triangular strain diagram shown in Fig. 6, the slip of the reinforcing bar slip can be expressed as s db fs slip = ------------8u (8)

The rotation at the end of the column due to slip of the reinforcing bars slip is given by the ratio of slip to the distance from the reinforcement to the neutral axis, c. Using Eq. (8), and recognizing that (s/c) is equal to the curvature at the section, the lateral displacement of an equivalent cantilever column of length a due to slip of the reinforcement at first yield can be expressed as follows ad b f s first yield slip first yield = a slip first yield = ---------------------------------8u (9)

where Av is the effective shear area of the column cross section (5/6 of the gross-section area of a rectangular column and 85% of the gross area of a circular column). The expected effects of concrete cracking suggest that the effective shear modulus should decrease as a function of the nominal principle tensile stress. Nonetheless, for application in engineering practice, the effective shear modulus Geff can be approximated as one half the elastic value for all levels of deformation. This value of the effective shear modulus was selected to optimize the statistics for the effective stiffness model developed below (Eq. (12)).

As shown in Fig. 1, the yield displacement (and each of its components) is defined as the displacement at an effective force of F0.004; hence, slip from Eq. (4) can be determined by multiplying Eq. (9) by the ratio F0.004 / Ffirst yield. Noting from Fig. 4 that y = first yield (M0.004/Mfirst yield) = first yield(F0.004/ Ffirst yield), the following expression for slip is derived ad b f s y slip = ----------------8u (10)

Fig. 6Deformations due to bar slip. 480

The average bond stress values recommended in the literature for elastic response range from u = 0.5 f c to 1.0 f c MPa (u = 6 f c to 12 f c psi) (Otani and Sozen 1972; ACI Committee 408 1979; Alsiwat and Saatcioglu 1992; Sozen et al. 1992; Lehman and Moehle 1998). For the purpose of this study, the average bond stress was taken as u = 0.8 f c MPa (u = 9.6 f c psi). At first yield of the column, the stress in the tension reinforcement, fs , used in Eq. (9) and (10) will vary depending on the axial load on the column. For columns with low axial loads, the tension reinforcement will yield; hence, fs can be taken as equal to the yield stress, fy. The stress in the tension reinforcement will decrease as the axial load on the column increases, reaching zero when the depth of the neutral axis is equal to the effective depth of the column. Consequently, it is expected that the displacement due to bar slip will increase with decreasing axial load. ACI Structural Journal/July-August 2009

To capture the effect of bar slip within the linear range of response, it is possible to include rotational springs at the ends of the column elements to directly model the additional flexibility from the slip of the longitudinal bars. The spring stiffness can be determined as 8u M 0.004 8u k slip = -------- ------------- = -------- EI flex d b fs y db fs (11)

axial loads to approximately 40% for stocky columns with low axial loads. The results shown in Fig. 7 also indicate that, except for stocky columns with high axial loads, shear deformations contribute less than 15% of the yield displacement for the columns in the database. Effective stiffness For engineering practice, it is convenient to use a single effective stiffness for a column element. Expressing EIeff calc as fraction of EIg and substituting Eq. (4) through (6) and (10) for y, Eq. (1) can be expressed as a function of nondimensional parameters EI eff calc ------------------- = ----------------------------------------------------------------------------------------EI g d b D f s f y 18 r v 2 D 2 E c 3 1 + -- ---- --- -- -- + ----- --- --- -------- - --8 D a f y u 5 D a G eff (12)

According to this approach and neglecting shear deformations, the effective stiffness of the column element, acting in series with the bond element, can be taken as EIflex from a momentcurvature analysis (Fig. 4). Contribution of components to total yield displacement As shown in Fig. 7, the contributions of flexure (Eq. (5)), shear (Eq. (6)), and bar slip (Eq. (10)) varied consistently with the axial-load ratio and aspect ratio. For both the rectangular and circular columns, the flexural mode of deformation contributed approximately 50 to 100% of the total deformation, depending on the level of axial load and the aspect ratio. The slip contribution ranged from 0% for columns with high

where =EIflex /EIg and rv is the radius of gyration of the column section in the direction of loading (r 2 = Ig/Av). For an v average bond stress value of u = 0.8 f c MPa (u = 9.6 f c psi), an effective shear modulus, Geff , equal to one half the elastic value, and using moment-curvature analysis to compute and fs /fy, Fig. 8 and Table 3 show the ratio of the measured effective stiffness to the effective stiffness determined using Eq. (12) for the column databases. EFFECTIVE STIFFNESS MODELS FOR PRACTICE For many practical situations, particularly those in which the column reinforcement has not yet been selected, it is preferable to use a version of Eq. (12) that does not require moment-curvature analysis. This section evaluates new models for effective stiffness that include the influence of bar slip. The models correspond to simplifications of Eq. (12), in which the results of moment-curvature analysis (that is, and fs /fy) are approximated and small terms are neglected. According to Eq. (12), the ratio of the effective stiffness to the gross-section stiffness is proportional to the normalized flexural rigidity, . This normalized flexural rigidity varies

Fig. 7Contribution of calculated deformation components as function of axial load and aspect ratio.

Fig. 8Comparison of calculated (Eq. (12)) and measured effective stiffnesses.

Table 3Statistics for ratio of measured to calculated effective stiffness for proposed models
Circular columns (108 specimens) Model Equation (12) Equation (16) Equation (17) Equation (18) with average bar size Minimum Maximum 0.50 0.63 0.64 0.57 1.69 1.80 1.76 1.59 Mean 1.04 1.04 1.04 1.02 Rectangular columns (221 specimens) Mean 0.97 0.93 0.92 0.95 Median 0.92 0.89 0.91 0.94 Coefficient of variation, % 26.6 26.6 26.9 25.5 Coefficient of Median variation, % Minimum Maximum 1.03 1.00 0.99 1.00 21.4 22.2 23.5 22.0 0.45 0.48 0.48 0.46 1.84 1.47 1.68 1.63

ACI Structural Journal/July-August 2009

481

primarily with the level of axial load, but also with the amount of longitudinal reinforcement. Assuming a linear stress-strain relationship for the concrete and steel, can be expressed in terms of the normalized initial strain due to the axial load, (P/AgEc)/o, and the relative stiffness of longitudinal reinforcement, n. The normalized flexural rigidity can be approximated as P Ag Ec approx = 0.2 + 1.3 ------------------- + n 1.0 o (13)

The form of Eq. (13) was derived by simplifying (using the binomial theorem) an analytical solution for the moment

of inertia of a generic cracked cross section with axial load. The coefficients in Eq. (13) were selected to provide the best possible estimate of the flexural rigidity determined from moment-curvature analysis (Fig. 4). Equation (13) provides a reliable substitute for moment-curvature analysis for a wide range of rectangular and circular columns. The ratio approx / (where is determined from a moment-curvature analysis) has a mean and coefficient of variation of 0.96 10.6% for rectangular columns and 1.04 9.5% for circular columns. For all of the 329 rectangular and circular columns, this ratio ranged from a minimum of 0.69 to a maximum of 1.23. If the reinforcement ratio has not yet been established, can be approximated by substituting an average value of n of 0.15 into Eq. (13), which results in the following relationship P Ag E approx = 0.35 + 1.3 ------------------c 1.0 o (14)

Fig. 9Approximation for steel stress as function of axial load.

Using Eq. (14), the mean and coefficient of variation for the rectangular and circular column databases are 0.95 14.4% and 1.04 20.7%, respectively. For the complete database, the ratios ranged from 0.59 to 1.63. Equation (14), whereas less accurate than Eq. (13), provides a good estimate of the normalized flexural rigidity for cases where the longitudinal reinforcement ratio is not yet known. As with , the ratio of the steel stress at column yield to the yield stress of the steel (fs /fy) can also be calculated with moment-curvature analysis. Alternately, it is possible to approximate this ratio by taking advantage of its dependence on the level of axial load. Figure 9, which compares the steel stress ratio determined based on moment-curvature analysis with the normalized initial strain (P/AgEc)/o for the database columns, indicates that the steel stress ratio can be approximated for both circular and rectangular columns as follows 4 10 P A g E c 0.0 f s f y approx = -- ----- ------------------- 1.0 3 3 o (15)

Equation (12) can be simplified by eliminating the third term in the denominator, which is related to shear deformations and tends to be small (refer to Fig. 7), by approximating the flexural rigidity with Eq. (13), and by approximating the steel stress ratio with Eq. (15). Taking advantage of these relationships, the need for momentcurvature analysis can be eliminated, and the effective stiffness can be approximated as
1.5 approx_Eq. 13 EI eff calc --------------------- = --------------------------------------------------------------------------------------- 1.0 and 0.2 EI g fs d b D 1 + 110 ---- --- ----------------------------------- - D a f y approx_Eq. 15

(16)

Fig. 10Comparison of calculated (Eq. (18)) and measured effective stiffnesses. 482

The two coefficients in Eq. (16) (1.5 and 110) were calibrated to compensate for the elimination of the shear term and to achieve a good match with the measured effective stiffness for the rectangular and circular column databases. As shown in Table 3, Eq. (16) provides similar levels of accuracy as Eq. (12), without requiring moment-curvature analysis. Using Eq. (14) (instead of Eq. (13)) to estimate , provides a model that does not require knowledge of the longitudinal reinforcement ratio without a significant decline in the model accuracy (refer to Table 3). ACI Structural Journal/July-August 2009

EI eff calc 1.5 approx_Eq. 14 --------------------- = --------------------------------------------------------------------------------------- 1.0 and 0.2 EI g fs d b D ----------------------------------- 1 + 110 ---- --- - D a f y approx_Eq. 15

(17)

The procedure can be simplified further by expressing the axial load as a fraction of Ag fc and recalibrating the constants, which results in the following relationship EI eff calc 0.45 + 2.5P A g fc ---------------------- = -------------------------------------------- 1.0 and 0.2 EI g db D - 1 + 110 ---- --- D a

coefficient of variation of 0.97 27% for rectangular columns and 1.04 21% for the circular columns. For the purpose of design, this paper simplifies the threecomponent model, with little loss of accuracy, by introducing an approximation to moment-curvature analysis, and by accounting for the effects of bar slip in terms of span-todepth ratio and axial-load ratio. According to this procedure, the effective column stiffness at yield can be estimated as EI eff calc 0.45 + 2.5P A g fc ---------------------- = -------------------------------------------- 1.0 and 0.2 EI g db D - 1 + 110 ---- --- D a (18)

(18)

To avoid the prior selection of the size of the longitudinal bars, the db /D term in Eq. (18) can be estimated prior to design for a given application. For bridge columns, the longitudinal bars tend to be well distributed around the column cross section, resulting in a smaller db/D term compared with building columns, where higher axial loads and architectural constraints tend to result in larger bar sizes relative to the column dimension. Based on the db/D values for the circular and rectangular column databases (Table 1), db /D can be taken as 1/25 for bridge columns and 1/18 for building columns. The form of Eq. (18) is consistent with the theoretical formulation of effective stiffness based on Eq. (12) and the summation of deformation components (Eq. (4) through (6) and 10). It provides a simple estimate of the effective stiffness without requiring a moment-curvature analysis or selection of the longitudinal reinforcement ratio. Figure 10 shows the ratio of the measured effective stiffness for the columns in the databases to the effective stiffness determined using Eq. (18) and the recommended average values for db/D. The lack of trends in Fig. 10 suggests that Eq. (18) properly accounts for the dependence of the effective stiffness on column axial load and aspect ratio. Furthermore, the data shows no bias with respect to the longitudinal reinforcement ratio. As shown in Table 3, Eq. (18) provides accuracy statistics that are consistent with those found for the much more complex Eq. (12). CONCLUSIONS The effective stiffnesses of 108 spiral-reinforced columns (with circular and octagonal cross sections) and 221 rectangular columns were estimated from data in the PEER Structural Performance Database (Berry et al. 2004). These data show that the normalized effective stiffness of the columns increases with increasing axial load, a trend that is reflected in many design guidelines and codes. The data also show that EIeff /EIg decreases with decreasing span-to-depth ratio, particularly for low axial loads. Seven existing models for column effective stiffness were evaluated using the column databases. The existing models were generally based on establishing an estimate for the flexural rigidity and ignored the influence of bar slip and shear deformations. All of the models tend to overestimate the measured effective stiffness and resulted in coefficients of variation ranging from 35 to 58%. This paper presented a three-component model, which explicitly combines the effects of flexure, bar slip, and shear components of deformation. This model reproduces the trends observed in the data and it provides an accurate estimate of column stiffness. For this model, the ratio of the measured effective stiffness to the calculated stiffness has a mean and ACI Structural Journal/July-August 2009

where the db /D can be approximated as 1/25 for bridge columns and 1/18 for building columns. This model, appropriate for implementation in design codes for concrete structures, is more accurate than existing models. Implementation of the proposed model resulted in a mean and coefficient of variation for the ratio of the measured effective stiffness to the calculated stiffness of 0.95 25.5% for rectangular columns and 1.02 22.0% for the circular columns. ACKNOWLEDGMENTS
This work was supported in part by the Earthquake Engineering Research Centers Program of the National Science Foundation, under Award Number EEC-9701568 through the Pacific Earthquake Engineering Research Center (PEER) and Natural Science and Engineering Research Council (NSERC) of Canada. The views expressed are those of the authors and not necessarily those of organizations cited here.

NOTATION
Ag Asl Av a aF0.004 aFfirst yield b c D db Ec Es EIeff meas EIflex EIg EIg,t e/D F0.004 Ffirst yield Fmax fc fs fy Geff l M0.004 Mfirst yield n P Pb Po Rc Rr rv Ts u = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = = gross cross-sectional area of column total area of longitudinal reinforcement effective shear area of column cross section shear span moment at maximum concrete compressive strain of 0.004 moment at first yield (Fig. 4) width of rectangular column section distance from tension reinforcement to neutral axis diameter (circular column) or column depth in direction of loading (rectangular column) nominal diameter of longitudinal bars concrete modulus of elasticity reinforcing steel modulus of elasticity measured effective stiffness (Eq. (1)) effective flexural stiffness (Fig. 4) gross bending stiffness gross transformed bending stiffness eccentricity ratio = M0.004 /PD effective force at maximum compressive strain of 0.004 (Fig. 1) effective lateral force at first yield (Fig. 1) maximum measured effective force concrete compressive strength (150 x 300 mm [6 x 12 in.] cylinders) longitudinal reinforcement steel stress at column fixed end longitudinal reinforcement yield stress effective shear modulus length over which bond stress acts (Fig. 6) aF0.004 = moment at maximum compressive strain at 0.004 aFfirst yield = moment at first yield (Fig. 4) modular ratio (Es/Ec) axial load (positive is compression) axial compression at balanced failure condition nominal axial load strength at zero eccentricity correlation coefficient for circular columns correlation coefficient for rectangular columns radius of gyration using shear area (rv2 = Ig /Av) tension force in one longitudinal bar (Fig. 6) constant bond stress

483

v slip flex shear slip

= = = = = =

slip first yield = y o s y first yield slip = = = = = = =

maximum nominal shear stress (Fmax /Ag) normalized effective flexural stiffness = EIflex/EIg slip of longitudinal reinforcing bar (Fig. 6) displacement at yield due to flexural deformations (Eq. (5)) displacement at yield due to shear deformations (Fig. (6)) displacement at yield due to bar-slip deformations (Eq. (10)) displacement at first yield due to bar-slip deformations (Eq. (9)) displacement at yield (refer to Fig. 1) nominal strain at which concrete is assumed to yield (0.002 in this study) longitudinal reinforcement steel strain at column fixed end curvature at yield (Fig. 4) curvature at first yield (Fig. 4) rotation at end of column due to slip of reinforcing bars (Fig. 6) longitudinal reinforcement ratio (Asl /Ag)

REFERENCES
ACI Committee 318, 2008, Building Code Requirements for Structural Concrete (ACI 318-08) and Commentary, American Concrete Institute, Farmington Hills, MI, 465 pp. ACI Committee 408, 1979, Suggested Development, Splice, and Standard Hook Provisions for Deformed Bars, Concrete International, V. 1, No. 7, July, pp. 44-46. Alsiwat, J. M., and Saatcioglu, M., 1992, Reinforcement Anchorage Slip under Monotonic Loading, Journal of Structural Engineering, ASCE, V. 118, No. 9, Sept., pp. 2421-2438. ASCE, 2000, Prestandard and Commentary for the Seismic Rehabilitation of Buildings, FEMA 356, Federal Emergency Management Agency, Washington, DC, Nov. ASCE, 2007a, Seismic Rehabilitation of Existing Buildings, ASCE/SEI 41, American Society of Civil Engineers, Reston, VA. ASCE, 2007b, Seismic Rehabilitation of Existing Buildings, ASCE/SEI 41, Supplement 1, American Society of Civil Engineers, Reston, VA. Benzoni, G.; Ohtaki, T.; Priestley, M. J. N.; and Seible, F., 1996, Seismic Performance of Circular Reinforced Concrete Columns under Varying Axial Load, SSRP 96/04, Structural Systems Research, University of California-San Diego, La Jolla, CA. Berry, M. P., and Eberhard, M. O., 2007, Performance Modeling Strategies

for Modern Reinforced Concrete Bridge Columns, PEER-2007/07, Pacific Earthquake Engineering Research Center, University of CaliforniaBerkeley, Berkeley, CA, 213 pp. Berry, M. P.; Parrish, M.; and Eberhard, M. O., 2004, PEER Structural Performance Database Users Manual, Pacific Earthquake Engineering Research Center, University of California-Berkeley, Berkeley, CA, www.ce.washington.edu/~peera1. Elwood, K. J.; Matamoros, A.; Wallace, J. W.; Lehman, D. E.; Heintz, J. A.; Mitchell, A.; Moore, M. A.; Valley, M. T.; Lowes, L. N.; Comartin, C.; and Moehle, J. P., 2007, Update of ASCE/SEI 41 Concrete Provisions, Earthquake Spectra, Earthquake Engineering Research Institute, V. 23, No. 3, Aug., pp. 493-523. Khuntia, M., and Ghosh, S. K., 2004, Flexural Stiffness of Reinforced Concrete Columns and Beams: Analytical Approach, ACI Structural Journal, V. 101, No. 3, May-June, pp. 351-363. Lehman, D. E., and Moehle, J. P., 1998, Seismic Performance of WellConfined Concrete Bridge Columns, PEER-1998/01, Pacific Earthquake Engineering Research Center, University of California-Berkeley, Berkeley, CA, 316 pp. Mander, J. B.; Priestley, M. J. N.; and Park, R., 1988, Theoretical Stress-Strain Model for Confined Concrete, Journal of Structural Engineering, V. 114, No. 8, pp. 1804-1826. Mehanny, S. S. F.; Kuramoto, H.; and Deierlein, G. G., 2001, Stiffness Modeling of RC Beam-Columns for Frame Analysis, ACI Structural Journal, V. 98, No. 2, Mar.-Apr., pp. 215-225. NZS 3101:2006, 2006, Commentary on the Concrete Design Standard, Part 2, Standards Association of New Zealand, Wellington, New Zealand. Otani, S., and Sozen, M. A., 1972, Behavior of Multistory Reinforced Concrete Frames during Earthquakes, Structural Research Series No. 392, University of Illinois at Urbana-Champaign, Urbana, IL, 551 pp. Paulay, T., and Priestley, M. J. N., 1992, Seismic Design of Reinforced Concrete and Masonry Buildings, John Wiley & Sons Inc., New York. Priestley, M. J. N.; Ranzo, G,; Benzoni, G.; and Kowalsky, M. J., 1996, Yield Displacement of Circular Bridge Columns, Caltrans Seismic Research Workshop, California Department of Transportation, Sacramento, CA, 12 pp. Sozen, M. A., 1974, Hysteresis in Structural Elements, Applied Mechanics in Earthquake Engineering, V. 8, pp. 63-98. Sozen, M. A.; Monteiro, P.; Moehle, J. P.; and Tang, H. T., 1992, Effects of Cracking and Age on Stiffness of Reinforced Concrete Walls Resisting In-Plane Shear, Proceedings of the 4th Symposium on Current Issues Related to Nuclear Power Plant Structures, Equipment and Piping, Orlando, FL.

484

ACI Structural Journal/July-August 2009

Potrebbero piacerti anche