Sei sulla pagina 1di 94

Assessment of current Colombian design practice for

reinforced concrete wall buildings

A dissertation submitted to the


ECOLE POLYTECHNIQUE FÉDÉRALE DE LAUSANNE
for the degree of
Master of Sciences

Presented by
Campiche Matteo

Supervised by
Project directors: Prof. Katrin Beyer, Ecole Polytechnique Fédérale de Lausanne, Switzerland
Prof. Carlos Andrés Blandón Uribe, Escuela de Ingeniería de Antioquia, Colombia
Prof. Ricardo León Bonett Díaz, Universidad de Medellín, Colombia
Senior researcher: Dr. João Pacheco de Almeida,
Ecole Polytechnique Fédérale de Lausanne, Switzerland

Spring 2014
The dissertation entitled “Assessment of current Colombian design practice for reinforced concrete wall build-
ings”, by Matteo Campiche, has been approved in partial fulfilment of the requirements for the Master Degree
in Civil Engineering.

Name of Reviewer 1 ____________________

Name of Reviewer 2 ____________________


Colombian RC wall buildings assessment – Abstract

ABSTRACT

In an attempt to reduce costs and construction time, a new common building trend has emerged in Colombia,
based on the use of particularly thin reinforced concrete walls with low reinforcement ratios. They can be
subjected to relatively large axial load ratios, depict low to no confinement, and often do not include boundary
elements. The abovementioned features have raised concerns amongst some design engineers, researchers,
and government officials, regarding their expected seismic performance. The objective of the present document
is to contribute with a first tentative answer to such understandable worries. To do so, an initial evaluation of
the current Colombian design and assessment code has been carried out, which identified room for future
improvements. Preliminary analyses on a set of recent Colombian buildings were then performed and their
results used to select a reference building showing typical features of the emerging construction trend. The
seismic evaluation of this building followed, according to two different Nonlinear Static Analysis methods. The
nonlinear properties of the models required to run the analyses were calibrated based on experimental data
collected from a recent laboratory test on a Colombian-type wall similar to those of the reference building.
The major conclusion of the analyses is that the selected reference building provides sufficient ductility to
sustain the design earthquake demands. Although reassuring and suggestive, this conclusion cannot obviously
be generalized to the Colombian building stock as further studies on other buildings would be required. It
should also be noted that the potential effects of wall out-of-plane instability and biaxial loading on the dis-
placement capacity were not directly considered in this study.

i
Colombian RC wall buildings assessment – Acknowledgments

ACKNOWLEDGMENTS

The author expresses his deepest gratitude to João Pacheco de Almeida, Katrin Beyer, Carlos Blandón, and
Ricardo Bonett for their guidance throughout the thesis. In particular, special thanks are due to João Pacheco
de Almeida for his valuable help and advices during the writing process of the present report.
The author also wishes to thank Jeyson Alzate, Sebastian Villaraga, Manuel Alejandro Restrepo and Juan
Manuel Maya all the preparatory work when collecting the information on the studied buildings, and David
Hernández, Luis Angel Morales, and Daniel Campiño for their assistance during the preliminary analysis.

ii
Colombian RC wall buildings assessment – Table of Contents

TABLE OF CONTENTS

ABSTRACT ....................................................................................................................................................... i
ACKNOWLEDGMENTS ................................................................................................................................. ii
TABLE OF CONTENTS ................................................................................................................................. iii
LIST OF FIGURES ........................................................................................................................................... v
LIST OF TABLES ........................................................................................................................................... vi
LIST OF SYMBOLS ....................................................................................................................................... vii
LIST OF EQUATIONS ................................................................................................................................... xii
1 INTRODUCTION ..................................................................................................................................... 1
1.1 Motivation ......................................................................................................................................... 1
1.2 Objective of the Master Project ......................................................................................................... 1
1.3 Scope of the Pre-Study ...................................................................................................................... 1
1.4 Scope of the Master Thesis ................................................................................................................ 2
2 COLOMBIAN DESIGN AND ASSESSMENT CODE: REVIEW AND COMPARISON WITH OTHER
CODES, WITH DESIGN PROVISIONS AND PUSHOVER ANALYSIS ..................................................... 3
2.1 General Design and Evaluation Procedure and Philosophy .............................................................. 3
2.1.1 Design procedure for new buildings .......................................................................................... 3
2.1.2 Evaluation, extension and rehabilitation procedure for existing buildings.............................. 12
2.2 Specific Requirements for Walls and their Seismic Design ............................................................ 17
2.2.1 General requirements for walls................................................................................................ 17
2.2.2 Specific requirements for seismic resistance of walls ............................................................. 19
2.3 Guidelines for the Pushover Analysis ............................................................................................. 23
2.4 Final Comments on the NSR-10 ...................................................................................................... 25
3 SELECTION AND DESCRIPTION OF THE REFERENCE BUILDING ............................................ 28
3.1 Selection of the Reference Building ................................................................................................ 28
3.1.1 Desired outcomes of the detailed analysis and selection criteria............................................. 28
3.1.2 Preliminary analysis procedure................................................................................................ 29
3.1.3 Final comments on the reference building selection ............................................................... 32
3.2 Reference Building Description and Preliminary Analysis Results ................................................ 32
3.2.1 Denomination convention........................................................................................................ 32
3.2.2 Description of the reference building ...................................................................................... 32
3.2.3 Preliminary analysis of the reference building ........................................................................ 36
3.3 Summary of the Reference Building Choice ................................................................................... 37
4 NUMERICAL MODELLING OF SINGLE WALL: VALIDATION WITH THE EXPERIMENTAL
RESULTS ........................................................................................................................................................ 38
4.1 Experimental Tests on a Thin RC Wall Representative of Colombian Construction Practice ........ 38
4.1.1 Geometry and reinforcement ................................................................................................... 38
iii
Colombian RC wall buildings assessment – Table of Contents

4.1.2 Material properties (MP) ......................................................................................................... 40


4.1.3 Loading .................................................................................................................................... 42
4.2 Modelling of the Wall with Two Distinct Finite Element Approaches ........................................... 42
4.2.1 VecTor2 Model........................................................................................................................ 43
4.2.2 SeismoStruct Model ................................................................................................................ 45
4.3 Comparison between Experimental and Numerical Response ........................................................ 48
4.3.1 Global level response............................................................................................................... 48
4.3.2 Local level response ................................................................................................................ 49
4.4 Conclusions on the Validity of the Numerical Modelling Approaches ........................................... 51
5 NUMERICAL MODELLING OF REFERENCE BUILDING: DESCRIPTION, ANALYSIS AND
RESULTS ........................................................................................................................................................ 52
5.1 Description of the Reference Building Model ................................................................................. 52
5.1.1 Geometry of the model ............................................................................................................ 52
5.1.2 Selected material behaviour models ........................................................................................ 53
5.1.3 Member modelling, floor constraints and other restraints ....................................................... 53
5.1.4 Mass attribution ....................................................................................................................... 55
5.2 Modal Analysis and Comparison with Preliminary Estimation of the Fundamental Period ........... 56
5.3 Nonlinear Static Analysis of the reference building ........................................................................ 58
5.3.1 Definition of the failure criterion............................................................................................. 58
5.3.2 Pushover analysis and application of the failure criterion ....................................................... 61
5.3.3 Behaviour of the reference building under design seismic loading ......................................... 62
5.4 Comparison of the of the Nonlinear Static Analyses, Conclusions, and Comments on the Validity of
the Preliminary Analysis ............................................................................................................................. 70
6 SUMMARY OF THE THESIS AND CONCLUSIONS ......................................................................... 71
REFERENCES ................................................................................................................................................ 73
APPENDIX A: LIST OF MODEL FILES .................................................................................................... A.1

iv
Colombian RC wall buildings assessment – List of Figures

LIST OF FIGURES

Figure 1.1 – Example of modern Colombian housing buildings (pictures: Prof. R. Bonett) ............................ 1
Figure 2.1 – Geographical zoning of the seismic hazard, (a) Aa (b) Av (c) seismic hazard zone [2] ............ 5
Figure 2.2 – Synoptic view of the procedure to determine the seismic hazard [2] ........................................... 6
Figure 2.3 – Seismic design categories [2] ........................................................................................................ 7
Figure 2.4 – Synoptic view of the analysis procedure [2] ................................................................................. 8
Figure 2.5 – Synoptic view of the procedure to determine the equivalent lateral seismic loads [2] ................. 9
Figure 2.6 – Synoptic view of the structural analysis under seismic loading and displacements evaluation [2]
......................................................................................................................................................................... 10
Figure 2.7 – Inter-story drift [2] ...................................................................................................................... 11
Figure 2.8 – Geographical zoning of the seismic hazard for limited safety level: Ae [2]................................ 14
Figure 2.9 – Pushover analysis procedure schematic representation............................................................... 24
Figure 3.1 – Representation of the preliminary analysis procedure ................................................................ 29
Figure 3.2 – Elastic displacement demand spectrum [2] ................................................................................. 31
Figure 3.3 – Floor plan .................................................................................................................................... 33
Figure 3.4 – Elevation of wall 1 ...................................................................................................................... 34
Figure 3.5 – Elastic displacement response spectra......................................................................................... 35
Figure 4.1 – 3D rendering of TW1 .................................................................................................................. 38
Figure 4.2 – TW1 geometry (including position of the centroid), dimensions in [mm].................................. 39
Figure 4.3 – Reinforcement layout .................................................................................................................. 40
Figure 4.4 – Stress-strain curves of the tensile test on the Ø6 and Ø16 rebars ............................................... 41
Figure 4.5 – Schematic representation of the loading process ........................................................................ 42
Figure 4.6 – Loading direction convention ..................................................................................................... 42
Figure 4.7 – VecTor2 model geometry and material numbering .................................................................... 43
Figure 4.8 – Stress-strain curves for steel rebars for VecTor2 model and selected experimental results ....... 45
Figure 4.9 – SeismoStruct Model geometry and loads .................................................................................... 46
Figure 4.10 – Stress-strain relationship of the concrete in the SeismoStruct model ....................................... 46
Figure 4.11 – Elements selected for the parametric analysis to define the confinement factor ...................... 47
Figure 4.12 – Stress-strain evolution for the concrete of the bottom 4 elements of the free extremity of the web
in the VecTor2 analysis when pushing in the positive direction ..................................................................... 47
Figure 4.13 – Stress-strain relationship of the rebars in the SeismoStruct model (a) Ø6 (b) Ø16 .................. 48
Figure 4.14 – Base shear vs top horizontal displacement: Comparison between numerical models (VecTor2
and SeismoStruct) and experimental results .................................................................................................... 49
Figure 4.15 – Vertical stress at the base vs top horizontal displacement comparison, flange side ................. 50
Figure 4.16 – Vertical stress at the base vs top horizontal displacement comparison, web side ..................... 50
Figure 5.1 – Reference Building SeismoSrtuct model .................................................................................... 52
Figure 5.2 – Link pattern used to simulate the slab effect ............................................................................... 54
Figure 5.3 – Influence zone of each walls ....................................................................................................... 54
Figure 5.4 – Deformed shapes of the 5 first vibration modes of the reference building model ...................... 56
Figure 5.5 – First and second vibration modes of a clamped beam in a two-dimensional space [36] ............ 57
Figure 5.6 – Global response of the reference building................................................................................... 62
Figure 5.7 – EPSH1 and EPSH2 idealizations of the building response in direction X+ ............................... 64
Figure 5.8 – EPSH1 and EPSH2 idealizations of the building response in direction Y+ ............................... 64
Figure 5.9 – EPP1 and EPP2 idealizations of the building response in direction X+ ..................................... 67
Figure 5.10 – EPP1 and EPP2 idealizations of the building response in direction Y+ ................................... 67
Figure 5.11 – Acceleration-Displacement Response Spectrum of the equivalent SDOF in direction X+ ...... 68
Figure 5.12 – Acceleration-Displacement Response Spectrum of the equivalent SDOF in direction Y+ ...... 69

v
Colombian RC wall buildings assessment – List of Tables

LIST OF TABLES

Table 2.1 – Synoptic view of the minimum energy dissipation capacity in relation to the seismic hazard level
........................................................................................................................................................................... 6
Table 2.2 – Synoptic representation of the structural system classification [2] ................................................ 7
Table 2.3 – Values of ϕc and ϕe ....................................................................................................................... 16
Table 3.1 –Spectral coefficients for damage control and design earthquakes [26] ......................................... 35
Table 3.2 – Period and displacement demand estimations for the reference building .................................... 36
Table 3.3 –Displacement ductility demand evaluation for the major walls of the reference building ............ 37
Table 4.1 - Reinforcement ratios of TW1........................................................................................................ 39
Table 4.2 - Uniaxial co mpression test results ................................................................................................. 40
Table 4.3 – Double Punch test results ............................................................................................................. 41
Table 4.4 - Average material properties of tested Ø6 and Ø16 rebars ............................................................ 41
Table 4.5 - List of material behaviour model used for the VecTor2 analysis ................................................. 44
Table 4.6 – Concrete material properties for the VecTor2 model ................................................................... 44
Table 4.7 - Steel rebar material properties for the VecTor2 model (with and without yield plateau) ............. 44
Table 5.1 - Mass attribution for each wall modelled in SeismoStruct as a separate element .......................... 55
Table 5.2 – Effective modal masses of the 5 first vibration modes of the reference building model.............. 56
Table 5.3 – Comparison of the fundamental vibration modes estimations with the eigenvalues analysis ...... 57
Table 5.4 – Calculation of the drift limit for the walls of the reference building (direction X) ...................... 60
Table 5.5 – Calculation of the drift limit for the walls of the reference building (direction Y) ...................... 60
Table 5.6 – Top lateral displacement at SD and NC limit states of the reference building for each direction 61
Table 5.7 – Detailed results of the Displacement Coefficient Method analysis .............................................. 65
Table 5.8 – Detailed results of the Capacity Spectrum Method analysis ........................................................ 69

vi
Colombian RC wall buildings assessment – List of Symbols

LIST OF SYMBOLS

Lowercase Latin Letters


as,max Maximum soil acceleration
bc Transversal dimension of the boundary element
bf Flange width
bw Minimum thickness of the element in the direction perpendicular to the shear load
c Distance from extreme compression fibre to neutral axis
d Structural depth of the element
dh Horizontal spacing of rebars
dv Vertical spacing of rebars
ds Minimum dimension of the side element OR cylinder diameter
do Minimum dimension of the side element
dp Diameter of punch
f c’ Specified concrete cylinder compressive strength
fcc (Confined) concrete peak stress
fs Steel rebar stress
ft' Concrete tensile strength
fu Ultimate tensile strength of rebars
fum Actual tensile strength of rebars
fy Specified yield stress of rebars
fym Actual yield stress of rebars
fyt Specified yield stress of the stirrups
g Earth standard acceleration
ios Overstress index
if Flexibility index
k Effective length factor OR adaptation factor for shear distribution on the building’s height
kc Concrete confinement factor
hi Altitude of story i (from the base) OR altitude of node i (from the base)
hw Wall height
hwi Total height of wall i
hx Altitude of story x (from the base) OR maximal centre-to-centre spacing between two rows
of confinement rebars
hpi Height of story (for story i)
hs Test unit height (cylinder, wall, etc)
hw Total wall height
lc Wall length
lw Wall length (in the considered direction)
lwi Horizontal length of wall i in the considered direction
lo Necessary penetration length of special transverse reinforcement
mi, mx Mass of story i or x OR nodal mass of node i
m* Modal mass
nw Number of walls resisting in the considered direction (for one building)
s, sx, sy Centre-to-centre rebar spacing (respectively in general, in direction x, and in direction y)
so Centre-to-centre longitudinal spacing of the confinement rebars
t Time OR element thickness
tw Wall thickness

Uppercase Latin Letters


Aa Acceleration factor
AB Story area of the building
Ac Gross area of concrete
Acv Gross area of concrete section bounded by web thickness and length of section in the direc-
tion of shear load considered
vii
Colombian RC wall buildings assessment – List of Symbols

Ae Acceleration factor for limited safety level design


Ash Total area of the stirrups (on a perpendicular section of height so)
Ast Total area of passive longitudinal reinforcement
At Acceleration in function of time
Av Velocity factor
Awi Shear area of wall i section in the considered direction
Ci Coefficient to calculate the yielding curvature with simplified equations
Cvx Coefficient of vertical repartition of the shear load for story x
C0 Modification factor to link the displacement of the control point with the displacement of
the equivalent SDOF (Pushover analysis according to the NSR-10 and FEMA-356)
C1 Modification factor to take into account the influence of the inelastic deformations in the
system response (Pushover analysis according to the NSR-10)
D Effects (internal loads, deflections) due to dead loads
D* Displacement of the equivalent SDOF
Dt Building top displacement
DSD Top displacement of the building corresponding to the Significant Damage limit state
E Reduced design seismic internal loads
Ec Concrete Young’s modulus
Es Steel rebars Young’s modulus
F Effects (internal loads, deflections) due to fluids with well-defined densities and controlla-
ble maximal depths
F* Force on the equivalent SDOF
Fa Site effect factor for acceleration
Fix, Fiy Equivalent lateral load of story i in direction x, respectively y
Fs Effects (internal loads, deflections) due to seismic design load (before reduction)
Fv Site effect factor for velocity
Fx Equivalent horizontal load of story x
G Effects (internal loads, deflections) due to hail
H Effects (internal loads, deflections) due to soil and ground water pressure
He Effective height of the equivalent SDOF
Hn Total height of the building
HRB Height of the reference building
HRB,s Height of the scaled reference building
I Importance factor
L Effects (internal loads, deflections) due to live loads
Le Effects (internal loads, deflections) due to stagnant water in or on the structure
Lr Effects (internal loads, deflections) of roof live loads
Ls Shear span (M/V)
LCSLS,i Load combination i for the serviceability limit state
LCULS,i Load combination i for the ultimate limit state
M Total mass of the building OR Flexural moment
𝑀̅ Moment profile over the height due to the unit force
Mu Design moment
N Number of stories in the building OR axial load
Nef Effective (design) strength of an element in an existing structure (axial, shear, flexion, etc)
Nex Existing strength of an element in an existing structure (axial, shear, flexion, etc)
Nu Design axial load
Pn Nominal axial strength of cross section
Q Quantity in the global system
Q* Quantity in the equivalent SDOF
R Energy dissipation factor (NSR-10)
R’ Energy dissipation factor for existing structures (NSR-10)
Ro Initial energy dissipation factor (NSR-10)
Rμ Reduction factor due to ductility (CSM)
Rd Ductility factor (Appendix A-3 of the NSR-10, and CSM)
S Soil profile type
Sa (Inelastic) Acceleration spectrum

viii
Colombian RC wall buildings assessment – List of Symbols

Sae (Elastic) Acceleration spectrum


Sae,max Plateau value of the elastic acceleration spectrum
Say Spectral acceleration at yielding point
Sd (Inelastic) Displacement demand (spectrum)
Sde (Elastic) Displacement demand (spectrum)
Sy Yield strength of the connection
T (Fundamental) period
Te Effective fundamental vibration period of the structure
To Period corresponding to the beginning of the acceleration spectrum plateau
To’ Period corresponding to the beginning of the constant ductility factor (Rd)
TC Period corresponding to the end of the acceleration spectrum plateau
TL Period corresponding to the beginning of the displacement spectrum plateau
T1 Tangent initial (fundamental) period of the structure, calculated from pushover analysis
TI, TII, TIII Estimations levels I, II and III for the building fundamental vibration period
V Base shear
Vd Design shear load
Ve Design shear load corresponding to the development of flexural resistance
Vj Total lateral load at step j of the pushover analysis
Vn Nominal shear strength of the section
Vs Shear load at the building base OR nominal shear strength provided by shear reinforcement
Vu Design shear sectional load OR ultimate shear strength
Vy Effective (idealized) yield base shear
VRd Design shear strength
VRd,c Design shear strength provided by the concrete without shear reinforcement
VRd,max Design maximum shear strength, defined by crushing of the concrete struts
VRd,s Design shear strength provided by shear reinforcement

Lowercase Greek Letters


α Factor for the decreasing part of the acceleration spectrum (microzoning campaign)
αc Factor for shear resistance depending on aspect ratio
γel Factor equal to 1.5 for primary and 1.0 for secondary seismic elements (EN 1998-3)
γM Material resistance safety factor (SIA, EN)
γSF Safety factor
δ Displacement of control point
δi Displacement of story i OR displacement of point i
δj Lateral displacement of the control point at step j of the pushover analysis
δT Target displacement of the control point
δu Design displacement OR ultimate displacement
δy Effective (idealized) yield displacement of the control point
δxi, δyi Displacements (in direction x or y) of story i
δNC Relative displacement of the first floor corresponding to the limit state of near collapse (SD,
EN)
δSD Relative displacement of the first floor corresponding to the limit state of significant damage
(SD, EN)
εc Strain in concrete
εsh Strain-hardening strain of rebars
εs Steel rebar strain
εt Strain in the rebars in tension
εu Ultimate strain of rebars, defined as strain corresponding to the tensile strength of rebars
εv Vertical strain
εy Yielding strain of rebars
εo Unconfined uniaxial concrete cylinder strain at fc’
θNC Chord rotation corresponding to the limit state of Near Collapse (NC)
θSD Chord rotation corresponding to the limit state of Significant Damage (SD)
λ Modification factor for reduced mechanical properties of lightweight concrete
μΔ Displacement ductility
ν Poisson’s ratio OR Axial load ratio

ix
Colombian RC wall buildings assessment – List of Symbols

ρ Reinforcement ratio
ρd Reinforcement ratio of diagonal reinforcement
ρl Longitudinal reinforcement ratio
ρo Orthogonal reinforcement ratio (in the flange)
ρt Transversal reinforcement ratio
σc Stress in concrete
ϕ Section resistance safety factor (NSR-10, ACI 318-08, NEC-11, E.60)
ϕa Energy dissipation reduction factor due to irregularities in elevation
ϕc Resistance reduction factor for quality of original design and construction, for evaluation
purposes
ϕe Resistance reduction factor for state of conservation, for evaluation purposes
ϕh Curvature profile over the height
ϕl,min Minimum longitudinal rebar diameter
ϕp Energy dissipation reduction factor due to plane irregularities
ϕr Energy dissipation reduction factor due to lack of redundancy of the structural system
ϕy Yielding curvature
ϕpl Plastic curvature
ϕo Confinement rebar’s diameter
ω Mechanical reinforcement ratio of the tension longitudinal reinforcement
ω’ Mechanical reinforcement ratio of the compression longitudinal reinforcement

Uppercase Greek Letters


Γ modal participation factor
Δp Plastic displacement
Δy Yield displacement
ΔNC Inter-story drift corresponding to the limit state of near collapse (NC, EN)
ΔSD Inter-story drift corresponding to the limit state of significant damage (SD, EN)
Δadm Admissible inter-story drift
Δloaded Lateral displacement at the point of load application
Δmax,i Maximum inter-story drift of story i (for any point of the diaphragm)
Δtracked Lateral displacement at the control point
Φi Displacement shape (for node i)
Ωo Overstrength factor

Other Symbols
Ø Rebar diameter (in mm)
@ Rebar spacing

Abbreviations and Acronyms


ACI 318M-08 American Building Code Requirements for Structural Concrete, version 2008
ACI 318M-11 American Building Code Requirements for Structural Concrete, version 2011
ADRS Acceleration-Displacement Response Spectrum
ASCE 7 ASCE/SEI 7-07: American Minimum Design Loads for Buildings and Other Structures
CCC SR-84 Código Colombiano de Construcciones Sismo Resistentes: 1984 version of the Colombian
Design Code for Buildings
CSM Capacity Spectrum Method
DCM Displacement Coefficient Method
EN European Norm: Eurocode, European Construction Codes
DES Disipación Especial: Special Energy Dissipation Capacity Category (code NSR-10)
DMI Disipación Minima: Minimum Energy Dissipation Capacity Category (code NSR-10)
DMO Disipación Moderada: Moderate Energy Dissipation Capacity Category (code NSR-10)
DSFM Disturbed Stress Field Model
E.60 Norma Técnica de Edificación, Concreto Armado, Peruvian Construction Code for RC
ED Elastic Demand
EESD Earthquake Engineering and Structural Dynamics Laboratory, EPFL
EIA Escuela de Ingeniería de Antioquia
EPFL École Polytechnique Fédérale de Lausanne

x
Colombian RC wall buildings assessment – List of Symbols

EPP1 First elastic-prefectly plastic idealization of the capacity curve


EPP2 Second elastic-prefectly plastic idealization of the capacity curve
EPSH1 First elastic-plastic idealization of the capacity curve with post-yield strain-hardening
EPSH2 Second elastic-plastic idealization of the capacity curve with post-yield strain-hardening
FEM Finite Element Method
FEMA-356 Prestandard and Commentary for the Seismic Rehabilitation of Buildings
infrmFB Inelastic Force-Based Frame Element (in SeismoStruct)
IP Integration Point
LMC Laboratoire de Matériaux de Construction at EPFL
MCFT Modified Compression Field Theory
MDOF Multiple Degree of Freedom Sytem
NC Near Collapse limit state (EN 1998-3)
NEC-11 Norma Ecuatoriana de la Construcción, Ecuadorian Construction Code
NSA Nonlinear Static Analysis
NSR-98 Normas Colombianas de Diseño y Construcción Sismo Resistente: 1998 version of the Co-
lombian Design Code for Buildings
NSR-10 Reglamento Colombiano de Construcción Sismo Resistente: 2010 version of the Colombian
Design Code for Buildings
PP Performance point
RC Reinforced Concrete
SD Significant Damage limit state (EN 1998-3)
SDOF Single Degree Of Freedom System
SIA Schweizerischer Ingenieur- und Architektenverein: Swiss Construction Codes
SLS Serviceability limit state
SS SeismoStruct, Software used for static and dynamic non-linear analysis of framed structures
developed by SeismoSoft
TD Target displacement
TW1 Thin Wall 1, Wall tested in the laboratories of EESD at the EPFL
UdeM Universidad de Medellín
ULS Ultimate limit state
V2 VecTor2, Software developed by the laboratories of the University of Toronto
YP Yielding Point

xi
Colombian RC wall buildings assessment – List of Equations

LIST OF EQUATIONS

(2.1) ................................................................................................................................................................... 7
(2.2) ................................................................................................................................................................... 9
(2.3) ................................................................................................................................................................... 9
(2.4) ................................................................................................................................................................... 9
(2.5) ................................................................................................................................................................... 9
(2.6) ................................................................................................................................................................. 10
(2.7) ................................................................................................................................................................. 10
(2.8) ................................................................................................................................................................. 10
(2.9) ................................................................................................................................................................. 11
(2.10) ............................................................................................................................................................... 11
(2.11) ............................................................................................................................................................... 11
(2.12) ............................................................................................................................................................... 12
(2.13) ............................................................................................................................................................... 12
(2.14) ............................................................................................................................................................... 12
(2.15) ............................................................................................................................................................... 16
(2.16) ............................................................................................................................................................... 16
(2.17) ............................................................................................................................................................... 16
(2.18) ............................................................................................................................................................... 18
(2.19) ............................................................................................................................................................... 18
(2.20) ............................................................................................................................................................... 18
(2.21) ............................................................................................................................................................... 19
(2.22) ............................................................................................................................................................... 19
(2.23) ............................................................................................................................................................... 19
(2.24) ............................................................................................................................................................... 19
(2.25) ............................................................................................................................................................... 19
(2.26) ............................................................................................................................................................... 19
(2.27) ............................................................................................................................................................... 20
(2.28) ............................................................................................................................................................... 20
(2.29) ............................................................................................................................................................... 21
(2.30) ............................................................................................................................................................... 21
(2.31) ............................................................................................................................................................... 21
(2.32) ............................................................................................................................................................... 21
(2.33) ............................................................................................................................................................... 22
(2.34) ............................................................................................................................................................... 22
(2.35) ............................................................................................................................................................... 22
(2.36) ............................................................................................................................................................... 23
(2.37) ............................................................................................................................................................... 24
(2.38) ............................................................................................................................................................... 24
(2.39) ............................................................................................................................................................... 24
(2.40) ............................................................................................................................................................... 24
(2.41) ............................................................................................................................................................... 24
(2.42) ............................................................................................................................................................... 25
(2.43) ............................................................................................................................................................... 25
(3.1) ................................................................................................................................................................. 29
(3.2) ................................................................................................................................................................. 29
(3.3) ................................................................................................................................................................. 30
(3.4) ................................................................................................................................................................. 30
(3.5) ................................................................................................................................................................. 30
(3.6) ................................................................................................................................................................. 30
(3.7) ................................................................................................................................................................. 30
(3.8) ................................................................................................................................................................. 30
(3.9) ................................................................................................................................................................. 31
xii
Colombian RC wall buildings assessment – List of Equations

(3.10) ............................................................................................................................................................... 31
(3.11) ............................................................................................................................................................... 31
(3.12) ............................................................................................................................................................... 32
(3.13) ............................................................................................................................................................... 32
(3.14) ............................................................................................................................................................... 35
(3.15) ............................................................................................................................................................... 35
(4.1) ................................................................................................................................................................. 47
(5.1) ................................................................................................................................................................. 57
(5.2) ................................................................................................................................................................. 57
(5.3) ................................................................................................................................................................. 58
(5.4) ................................................................................................................................................................. 58
(5.5) ................................................................................................................................................................. 58
(5.6) ................................................................................................................................................................. 59
(5.7) ................................................................................................................................................................. 59
(5.8) ................................................................................................................................................................. 59
(5.9) ................................................................................................................................................................. 59
(5.10) ............................................................................................................................................................... 62
(5.11) ............................................................................................................................................................... 63
(5.12) ............................................................................................................................................................... 63
(5.13) ............................................................................................................................................................... 63
(5.14) ............................................................................................................................................................... 63
(5.15) ............................................................................................................................................................... 63
(5.16) ............................................................................................................................................................... 66
(5.17) ............................................................................................................................................................... 66
(5.18) ............................................................................................................................................................... 67
(5.19) ............................................................................................................................................................... 68
(5.20) ............................................................................................................................................................... 68
(5.21) ............................................................................................................................................................... 68
(5.22) ............................................................................................................................................................... 68
(5.23) ............................................................................................................................................................... 68

xiii
Colombian RC wall buildings assessment – Introduction

1 INTRODUCTION

Figure 1.1 – Example of modern Colombian housing buildings (pictures: Prof. R. Bonett)

1.1 Motivation
Over the last few years, due to a lack of residential housing and the relatively high cost of the land in several
metropolitan areas, a large number of mid to high rise reinforced concrete (RC) buildings has been constructed
in Colombia. The material and structural system of RC shear walls has proven to effectively sustain lateral
wind and seismic loading, providing adequate strength, stiffness, as well as deformation and energy dissipation
capacities.
In an attempt to reduce costs and construction time, and due to the relatively high price of materials in com-
parison with labour, a new trend has emerged in Latin-American countries; construction with particularly slen-
der walls and using low vertical and horizontal reinforcement ratios, often without boundary elements. This
lead to walls subjected to large axial load ratios, and low to non-existent confinement, hence compromising
the ability to sustain lateral loads, and the required deformation capacity. This new trend has been especially
noticeable in Colombia, Ecuador and Peru.

1.2 Objective of the Master Project


The present master project results from a cooperation between the Ecole Polytechnique Fédérale de Lausanne
(EPFL), the Escuela de Ingeniería de Antioquia (EIA), and the Universidad de Medellín (UdeM). It intends to
provide a review of the Colombian code provisions regarding general building design and evaluation, as well
as of the design of reinforced concrete walls. It also aims to supply detailed analyses and conclusions regarding
the in-plane behaviour of typical Colombian slender reinforced concrete walls. In order to assess their behav-
iour, the evaluation of a typical building, engineered in accordance with the Colombian standards will be
carried out using nonlinear static analysis (Capacity Spectrum Method, CSM).

1.3 Scope of the Pre-Study


The aim of the work (herein referred as pre-study) carried out by the student in preparation of the present thesis
was to gain sufficient theoretical background in topics related to thin shear RC walls with low reinforcement
ratios. The student also took part in the modelling of Thin Wall 1 (TW1), tested in the Earthquake Engineering
and Structural Dynamics (EESD) laboratory. This test unit, presented in chapter 4, showed features of typical
Colombian construction practices (high slenderness, low reinforcement ratio, absence of boundary elements).
1
Colombian RC wall buildings assessment – Introduction

This modelling aimed to define the loading process of the laboratory experiment, verify the validity of the
simulations carried out with the nonlinear membrane structural analysis software VecTor2 (V2), as well as to
get acquainted with the features of typical Colombian design practices. With the previous objectives in mind,
the following tasks have been carried out during the pre-study:
 Review of the scale effect and damage localization in compressed concrete and finite element models
(FEMs)
 Review of the models used for material behaviour in VecTor2 simulation:
o Concrete behaviour
o Reinforcement bars
 Modelling of a typical Colombian wall (TW1, Thin Wall 1):
o Description of the tested wall
o Description of the model (Geometry, used elements and reinforcement types)
o Model predictions and interpretation

1.4 Scope of the Master Thesis


After the Pre-Study have been completed, the core research of the master thesis, focused on the in-plane be-
haviour of typical Colombian walls will be carried out, according to the following layout:
In-Plane behaviour of Walls:
 Review of the Colombian construction code philosophy and important points regarding shear wall
design, and comparison with other codes (Peru, Ecuador, USA, Switzerland, Europe)
 Simplified evaluation of a set of existing building in order to select a representative building to be
modelled in a nonlinear analysis
 Verification of the validity of the selected modelling options by comparing experimental and numeri-
cal analysis results for TW1
 Evaluation of the seismic behaviour of the selected building by applying a displacement-based assess-
ment method

2
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

2 COLOMBIAN DESIGN AND ASSESSMENT CODE: REVIEW AND COMPARI-


SON WITH OTHER CODES, WALL DESIGN PROVISIONS AND PUSHOVER
ANALYSIS

The present chapter provides an overview of the Colombian code general philosophy with a focus on seismic
resistant walls and the special requirements of those load-bearing systems. Some comparisons with other in-
ternational construction codes (e.g. the Swiss SIA, the European EN, etc.) are made in order to situate the
Colombian code in the international codes context.
The first Colombian construction code for buildings, the Codigo Colombiano de construcciones Sismo Re-
sistentes, CCC SR-84, appeared in 1984. Before that, there was no national construction code and it was the
choice of the design engineer to select what procedure to apply. Mostly, engineers used the American and
sometimes German standards available at the time. The CCC SR-84 was then replaced 14 years later in 1998
by the Normas Colombianas de Diseño y Construcción Sismi Resistente, NSR-98 [1], which remained valid
until the introduction of the current design code, the Reglamento Colombiano de Construcción Sismo Re-
sistente, NSR-10 [2].
The NSR-10, as the vast majority of the south and central American codes, is largely based on the United
States codes; Chapters A (“General design and constructive requirements for earthquake resistant construc-
tions”) and B (“Loads”) of the NSR-10 are mainly based on the ASCE 7 (Minimum Design Loads for Buildings
and Other Structures [3]), whilst chapter C (“Structural concrete”) is an adaptation of the ACI 318M-08 (Build-
ing Code Requirements for Structural Concrete [4]). It is to be noted that the latest version of the ACI (ACI
318M-11 [5]), which consisted mainly in a reorganization of the previous code structure, has not been taken
into account in the current Colombian code.
Chapter 2.1 of the present thesis offers an overview of the general evaluation procedure and philosophy of the
NSR-10, while chapter 2.2 focuses on the specific requirements for walls and their seismic design, and chapter
2.3 presents the recommendations given in appendix A-3 of the NSR-10 for pushover analysis. Finally, chapter
2.4 gathers the final comments on the NSR-10.

2.1 General Design and Evaluation Procedure and Philosophy


The Colombian design code for buildings is mainly focused on seismic resistance, as reflected by its name
(Colombian Regulation of Seismic-Resistant Constructions, free translation) and its first chapter dedicated to
seismic loading.
The NSR-10 (chapter A) proposes a detailed structural design (and evaluation) procedure for new and existing
buildings. For the sake of clarity and readability of the present report, the procedure for the design of new
buildings is first presented in chapter 2.1.1, and the important differences for the evaluation of existing struc-
tures are then explained in chapter 2.1.2.

2.1.1 Design procedure for new buildings


2.1.1.1 Step D1: Preliminary design and coordination with other professionals
During this step, the structural design engineer will define the structural system and dimensions in order to
assess the mass, loads (dead, live, wind, seismic loads etc.). This needs to be done in collaboration with all
major protagonists of the construction process, in order to include decisive issues in the original design, such
as fire safety or potential future reassignments of building allocation.
2.1.1.2 Step D2: Evaluation of the structural loads (apart from seismic loading)
Once the final dimensions of all structural members are defined, all the loads arising from dead, live and wind
loads, imposed deformation (geotechnical, temperature, creep, shrinkage), etc, can be evaluated, as well as the
mass of the building and its distribution. It is to be noted that the code explicitly requires to include part (25%)

3
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

of the live loads in the building mass when proceeding to the modal analysis (and mass distribution) for ware-
houses only, without stating anything on the subject for other types of buildings. It seems counter-intuitive
because an important part of the live loads is actually permanent (e.g. furniture in a housing building).
2.1.1.3 Step D3: Determination of the seismic hazard level
Depending on the geographical location of the designed building, the design peak values of ground accelera-
tion and velocity factors are defined according to historical seismic activity and a certain return period, herein
taken as 475 years as in many other standards (e.g. the Swiss SIA 261 [6], the German DIN 4149 [7]). It is to
be noted here that the NSR-10 uses independent factors for the response of vibration modes with long and
short periods. Namely, the factor Aa, representative of acceleration, is used for short period vibration modes,
while Av, representative of velocity, is used for long period vibration modes. This separation is explained by
the fact that vibration modes with short and long periods are influenced by earthquakes having different
sources, hence the dissociation. The delimitation between short and long period is defined as TC, i.e. the period
corresponding to the end of the acceleration spectrum plateau.
Finally, depending on the maximum value between Aa and Av, a seismic hazard zone is attributed (low, inter-
mediate or high) that will later define the construction requirements for details, limitations for certain types of
structural systems (i.e. maximum height of buildings, maximum number of stories, Usage Group restriction),
etc.
The different zones for those three parameters can be seen in Figure 2.1. The detailed procedure is illustrated
in the preliminary analysis shown in chapter 3.1.2.1.2 of the present report.
2.1.1.4 Step D4: Determination of design spectrum and ground motions
For the standard analysis using the code spectra, the NSR-10 proposes to use the site amplification factors of
Fa and Fv depending on the values of Aa and Aa and on the profile of the first 30 m of soil on which the structure
will be standing. The soil is characterized by the mean shear waves speed, mean shear resistance, mean punch-
ing resistance, plasticity index and thickness of its layers.
The last factor to define the response spectra is the Usage Group (in European terminology: Importance Class)
of the building, ranging from I to IV to which corresponds an importance factor I varying from 1.00 to 1.50
representing an increase of the return period of the spectra. This Usage Group depends on the type of use of
the structures and the consequences of a failure/collapse and/or of an eventual downtime during the repair/re-
construction time.
The code also offers the possibility to use a family of accelerograms for dynamic analysis or to use data of
microzoning campaigns (generally providing design spectra), provided for both cases that the data is sufficient.
Figure 2.2 presents a synoptic representation of the procedure to compute the design seismic hazard.
In addition to the normal design spectrum, chapter A.12 of the NSR-10 sets a Damage control spectrum for
Usage Groups III and IV. This spectrum aims to ensure that constructions essential to the recovery of the
society after an earthquake do not suffer downtime for earthquakes of moderate intensities. The only verifica-
tion to be carried-out with this spectrum consists of a comparison of the drifts with flat-rate admissible drift
values. The damage control earthquake has a probability of occurrence of 80% on a period of 50 years, i.e. a
return period of 31.6 years.
It is to be noted that all the response spectra defined in the NSR-10 are elastic response spectra.

4
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

(a) (b)

(c)
Figure 2.1 – Geographical zoning of the seismic hazard, (a) Aa (b) Av (c) seismic hazard zone [2]

5
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

Figure 2.2 – Synoptic view of the procedure to determine the seismic hazard [2]
2.1.1.5 Step D5: Definition of the structural system and material characteristics
Following the same philosophy as the ASCE 7, the NSR-10 requires structural categorization depending on
the structural system bearing vertical and horizontal loads (bearing walls, frames, dual, combined, see Table
2.2), materials (steel, reinforced concrete, masonry), and energy dissipation capacity, see Figure 2.3.
However, while the ASCE 7 separates the buildings in 6 distinct seismic design categories (A, B, C, D, E and
F) depending on their energy dissipation capacity, the NSR-10 reduces those categories to a number of 3: DMI,
DMO and DES (see Table 2.1), which stand respectively for minimum, moderate and special energy dissipa-
tion capacity. Those categories, combined with a defined structural system, correspond to an assumed initial
energy dissipation factor Ro and over-strength factor Ωo.
Depending on the seismic hazard zone, the NSR-10 restricts (in terms of the maximum height, number of
stories or Usage Group) or forbids certain structural system types, and seismic design categories. For instance,
in the region of Medellín and Envigado, the seismic hazard is classified as intermediate. Hence, a combined
structural system composed of RC walls of category DMO for horizontal loads and RC frames of category
DMO for vertical loads is allowed up to heights of 72 m, according to table A.3-2 of the NSR-10, while a
system composed of RC bearing walls of category DMO for both loading directions is allowed up to heights
of 50 m according to table A.3-1.
Table 2.1 – Synoptic view of the minimum energy dissipation capacity in relation to the seismic hazard level
Seismic hazard level
Energy dissipation capacity
Low Intermediate High
Minimum DMI
Moderate DMO
Special DES

6
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

Table 2.2 – Synoptic representation of the structural system classification [2]

Figure 2.3 – Seismic design categories [2]


2.1.1.6 Step D6: Level of irregularity of the structure and analysis procedure
In order to take into account irregularities of the building that might lead to early collapse, the NSR-10 intro-
duces three factors of reduction of energy dissipation capacity, ϕa, ϕp and ϕr, illustrated in Figure 2.4. Those
factors represent respectively the irregularities in elevation, the in plan irregularities, and the lack of redun-
dancy of the structural system. The final energy dissipation factor is calculated with eq. (2.1).
𝑅 = 𝜙𝑎 ∙ 𝜙𝑝 ∙ 𝜙𝑟 ∙ 𝑅𝑜 (2.1)
where Ro is the initial energy dissipation factor prescribed by the NSR-10 depending on the structural system
category the energy dissipation category it is assigned to, ϕa, ϕp and ϕr are the irregularity reduction factors,
and R is the final design energy dissipation factor

7
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

Figure 2.4 – Synoptic view of the analysis procedure [2]


In this step, the design engineer must also define the analysis methodology. The NSR-10 allows 4 types of
analysis, as described below:
(a) Equivalent lateral force method (Chapter A.4 of the NSR-10), for the following types of buildings:
i. All buildings in zones of low seismic hazard zone
ii. All buildings of Usage Group I in zones of intermediate and low seismic hazard zone
iii. Regular buildings of 20 stories or less and no higher than 60 m from their base in any seismic
hazard zone, except for buildings on poor soil profiles having high fundamental vibration pe-
riods
iv. Irregular buildings with no more than 6 stories and not higher than 18 m
v. Flexible structures supported by structures of significantly larger stiffness
(b) Dynamic elastic analysis methods (Chapter A.5 of the NSR-10):
i. All structures that do not comply with the requirements for the application of the equivalent
lateral force method
(c) Dynamic inelastic analysis methods (Chapter A.5 of the NSR-10):
i. When, according to the design engineer, the structure may present a large non-dependable
variation in the energy dissipation capacity of its members that can only be assessed by this
method.
(d) Alternative analysis methods (quasi-static pushover, etc.), which must take into account the dynamic
and inelastic characteristics of the structure (Appendix A-3 of the NSR-10):
i. When the engineer wishes to assess the energy dissipation capacity in the inelastic range
and/or the maximum displacement capacity with this method. The use of this type of methods
is not compulsory, it is a voluntary choice of the engineer.
For the verification of structures of Usage Groups III and IV with the damage control earthquake, article
A.12.4.1 states that the structural analysis must be carried out with the equivalent lateral force method, or with
the dynamic analysis method, presented in chapter A.5 of the NSR-101.

1
The aim of the damage control spectrum being to maintain the structure in its elastic domain according to
article A.12.1.3, one can assume that the dynamic method mentioned here refers solely to the elastic dynamic
method of chapter A.5.
8
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

2.1.1.7 Step D7: Determination of the seismic inertial loads


In order to calculate the equivalent lateral load, the fundamental period of the structure must first be determined
by means of empirical formulae (shown in section 3.1.2 of the present report). Once the final design of the
structure is achieved, the period must be recalculated by a modal analysis or a deflection analysis under lateral
static loading. If the final period differs by more than 10% from the originally calculated one, the NSR-10
prescribes a repetition of the seismic analysis until convergence is reached.
It is interesting to note that the NSR-10, like the SIA 261 [6], proposes to only consider the fundamental period
but unlike the latter, does not impose an upper limit on the fundamental period for the application of this
method (2.0 s in the SIA 261). It rather introduces a different formula for the lateral load distribution, attrib-
uting larger loads to the highest stories when the period increases. The procedure proposed in the NSR-10 can
be observed in eqs. (2.2) to (2.5) as well as Figure 2.5. The procedure proposed in the SIA 261 is similar (with
another formulation), except that the k factor present in eq. (2.4) has a constant value of 1.0 s (in addition to
the validity condition of T ≤ 2.0 s).
𝑉𝑠 = 𝑆𝑎𝑒 ∙ 𝑔 ∙ 𝑀 (2.2)
𝐹𝑥 = 𝐶𝑣𝑥 𝑉𝑠 (2.3)
𝑚𝑥 ∙ ℎ𝑥𝑘
𝐶𝑣𝑥 = (2.4)
∑𝑛𝑖=1(𝑚𝑖 ∙ ℎ𝑖𝑘 )
1.0 for 𝑇 < 0.5 [s]
𝑘 = {0.75 + 0.5 ∙ 𝑇 for 0.5 [s] < 𝑇 < 2.5 [s] (2.5)
2.0 for 𝑇 > 2.5 [s]
where Vs is the seismic base shear, Sae stands for the elastic spectral acceleration, M is the total building mass,
including permanent loads (weight of permanent equipment, non-structural walls etc.), g represents the accel-
eration of gravity, mx and mi are the masses of stories x and i, hx and hi are the heights of stories x and i,
measured from the base of the structure, and finally T is the fundamental period of the building.
As noted above, the NSR-10 only requires to take into account a part (25%) of the live loads in the total
building mass for warehouses, leaving the engineer free to consider or not the live load in other cases.
For the dynamic analysis methods, the loads to be applied are the series of accelerograms, scaled to fit the
expectable seismic hazard characteristics for the region, considering site effects.

Figure 2.5 – Synoptic view of the procedure to determine the equivalent lateral seismic loads [2]

9
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

2.1.1.8 Step D8: Structural analysis under seismic loading


Once the design seismic loads have been defined in step D7, the structural analysis of the building can be
carried out according to the method defined in step D6 in order to determine the internal loads Fs (Figure 2.6)
due solely to seismic design loads.
2.1.1.9 Step D9: Evaluation of the horizontal displacements
The evaluation of the horizontal displacements (Figure 2.6) including P-Δ effects must be carried out taking
into account the eventual torsional effects, and using the properties of the elastic, uncracked section of the
lateral load bearing elements.
Alternatively, an assessment of the effective stiffness of elements as a percentage of the gross inertia is pro-
posed in article C.10.10.4.1(b) for different elements (columns, walls, beams, slabs etc.) to take cracking of
concrete into account.
The load combinations LCSLS,1, LCSLS,2, and LCSLS,3 used to estimate the horizontal displacement are given in
eqs. (2.6) to (2.8), according to articles B.2.3.1 and B.2.3.2.
𝐿𝐶𝑆𝐿𝑆,1 = 𝐷 + 𝐻 + 𝐹 + 𝐹𝑠 (2.6)
𝐿𝐶𝑆𝐿𝑆,2 = 𝐷 + 𝐻 + 𝐹 + 0.75𝐹𝑠 + 0.75𝐿 + 0.75(𝐿𝑟 or 𝐺 or 𝐿𝑒 ) (2.7)
𝐿𝐶𝑆𝐿𝑆,3 = 0.6𝐷 + 𝐹𝑠 + 𝐻 (2.8)
where Fs represents the effects of the design seismic load, D denotes the effects of the dead load of the struc-
ture, non-structural elements and permanent equipment, L symbolizes the effects of live loads, H depicts the
effects of soil and ground water pressure, F stands for the effects of fluids with well-defined densities and
controllable maximal depths, Lr represents the effects of roof live loads, G denotes the effect of hail, and Le
symbolizes the effects of stagnant water in or on the structure (for instance on the roof). It is to be noted that
the term "effects" stands for the deflections of elements in the context of the Serviceability Limit State (SLS).

Figure 2.6 – Synoptic view of the structural analysis under seismic loading and displacements evaluation [2]

10
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

Figure 2.7 – Inter-story drift [2]


2.1.1.10 Step D10: Verification of the inter-story drift
Once the computation of the horizontal displacements along the building height is completed, the NSR-10
requires verifying that the maximum inter-story drift for any point of the diaphragm Δmax,i (Figure 2.7), does
not exceed the admissible value Δadm of 1.00% (0.50% for fragile masonry constructions, according to the
errata published on March 19th, 2010 [8]). These rather strict limits are set to guarantee the integrity of fragile
non-structural elements (infill masonry for instance). If cracking was taken into account in the definition of
the structure stiffness and in the analysis, it is allowed to multiply the calculated drifts by a factor of 0.7 before
comparing them with the admissible values. However, the use of cracked sectional properties in the analysis
is judged to have larger influence on the estimation of displacements and therefore the indulgence granted for
this technique is considered not great enough to counteract the enhanced calculated displacements. Design
engineers hence usually prefer to use gross sections in their analyses.
For the drift verification of Usage Groups III and IV subjected to the damage control spectrum, the admissible
limits Δadm are respectively 0.40% for reinforced concrete, timber, steel, and reinforced masonry structures,
and 0.20% for plain masonry structures.
2.1.1.11 Step D11: Combination of the different loads and design of the structural elements
For what regards load combination, the NSR-10 proposes (for methods (a) and (b) of Step D6) to divide the
internal loads Fs coming from the seismic analysis by the same coefficient of energy dissipation capacity R
over the whole structure. This global energy dissipation factor is applied regardless of whether or not the
different elements actually enter the plastic regime, where actual energy dissipation occurs. These design load
combinations LCULS,1 and LCULS,2 are shown in eqs. (2.10) and (2.11). Note that only load combinations in-
cluding earthquake loading are presented here.
𝐸 = 𝐹𝑠 /𝑅 (2.9)
𝐿𝐶𝑈𝐿𝑆,1 = 1.2𝐷 + 1.0𝐸 + 1.0𝐿 (2.10)
𝐿𝐶𝑈𝐿𝑆,2 = 0.9𝐷 + 1.0𝐸 + 1.6𝐻 (2.11)
where E represents the reduced design seismic internal loads (including energy dissipation effects), D denotes
the effects of the dead load of the structure, non-structural elements and permanent equipment, L symbolizes
the effects of live loads, and H depicts the effects of soil and ground water pressure. Note that the term "effect"
denotes internal loads and loads in the context of Ultimate Limit State (ULS).
It is to be noted that the NSR-10 as well as the ACI 318M-08 uses the term factored (e.g. factored loads) in a
context where the European codes (e.g. the Swiss construction code SIA 260 [9] and the European Eurocode
EN 1990 [10]) would use the denomination design (e.g. design load). The European terminology (design loads)
will be used throughout the present report.
2.1.1.12 Step D12: Design of the structural elements
Depending on the energy dissipation category of the building (DMI, DMO and DES), the design of the differ-
ent structural elements (beams, walls, columns) must comply with different requirements, described in a dif-
ferent chapter or sub-chapter of the NSR-10 for each element type and material.
A significant difference between the force-based design of the European codes (EN 1990 [10], SIA 260 [9])
and the NSR-10 (as well as the ACI 318M-08 [4], the E.060 [11] from Peru, the NEC-11 [12] from Ecuador
and other South American codes), is the fact that the safety factor on the resistance is not separately applied
on each material as it is the case for the SIA, but it is dependent on the expected failure mode and is presented
as a reduction (multiplying) factor on the total section resistance. For instance, in the NSR-10 and
11
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

ACI 318M-08, with an element subjected to flexion, depending on the relative deformation on each side of the
section, the safety factor will vary from ϕ = 0.65, for an expected compression-controlled failure (concrete
reaches εc = 3‰ before the rebars reach εt = 2‰), to ϕ = 0.90, when the expected failure is tension-controlled
(concrete reaches εc = 3‰ after the rebars reach εt = 5‰), and will vary linearly between those two cases. Each
load transfer mode (shear/torsion, rebar anchoring, contact pressure on concrete, etc.) has then its specific
safety factor ϕ.
Another significant difference between the European and the American codes can be found in the way the
shear strength of beams (and shear walls) is considered for reinforced concrete. The approach proposed by the
SIA 262 [13] and the EN 1992 [14] (both for concrete structures) depends on the level of loading. If the load
is low enough to be sustained by the concrete only, the shear resistance of a section is determined by a formula
taking into account for the longitudinal reinforcement. Note that the exact formula to calculate the shear re-
sistance provided by the concrete is different in SIA 262 and EN 1992. If the load is such that shear reinforce-
ment is required, the total shear load is supposed to be transferred through cracks by shear reinforcement (when
no inclined compressed or tension chords are present), and a higher bound of the resistance is provided by the
crushing limit of the concrete struts. This kind of approach leads to eq. (2.12).
𝑉𝑅𝑑,𝑐 when 𝑉𝑑 ≤ 𝑉𝑅𝑑,𝑐
𝑉𝑅𝑑 = { (2.12)
𝑉𝑅𝑑,𝑠 ≤ 𝑉𝑅𝑑,𝑚𝑎𝑥 when 𝑉𝑑 > 𝑉𝑅𝑑,𝑐
where VRd is the design shear strength of the section, VRd,c represents the (design) shear strength provided by
the concrete without shear reinforcement (but indirectly considering the longitudinal reinforcement), Vd is the
design shear load, VRd,s stands for the design shear strength provided by shear reinforcement, and VRd,max is the
design maximum shear strength, defined by crushing of the concrete struts.
The NSR-10 and the ACI-318 provide a different model, based on the simple sum of the resistance provided
by concrete and shear reinforcement, as illustrated in eqs. (2.13) and (2.14).
𝜙𝑉𝑛 ≥ 𝑉𝑢 (2.13)
𝑉𝑛 = 𝑉𝑐 + 𝑉𝑠 (2.14)
where Vn is the nominal shear strength of the section, Vu the design shear load, ϕ the strength reduction factor
(in this case for shear resistance), Vc the nominal shear strength provided by concrete and Vs the nominal shear
strength provided by shear reinforcement.

2.1.2 Evaluation, extension and rehabilitation procedure for existing buildings


In Colombia, many of the existing buildings were built before 1984, when the first Colombian construction
code was introduced. This poses a major problem regarding the heterogeneity of the construction’s quality. To
deal with this reality, chapter A.10 of the NSR-10 provides a procedure to assess the existing buildings, defin-
ing a distinct procedure for structures designed under each successive version of the Colombian construction
codes (before the CCC SR-84 [15] of 1984, under the CCC SR-84, under the NSR-98 [1] of 1998, and under
the NSR-10 of 2010).
Chapter A.10 of the NSR-10 presented here offers guidelines to proceed with a vulnerability analysis, and
provides separate recommendations depending on the desired structural intervention (attached extension, ver-
tical extension, seismic rehabilitation, and retrofit after an earthquake). Those guidelines are meant to ensure
that existing structures are able to resist to small earthquakes without damages, median earthquakes without
structural damages, and major earthquakes without collapse.
2.1.2.1 Preliminary information
2.1.2.1.1 Step E1: Verification of the validity of the procedure
The first step of this procedure is to verify its validity. Indeed, this method is only valid for the evaluation of
existing structures, and not for new constructions. Amongst the different cases of application of the present
method, the NSR-10 mentions the following:
It must be applied for:
(a) Cases of a change of allocation of the building in a urban planning sense (e.g. residential to commercial
etc.), especially when there is a change for a superior Usage Group (as defined in chapter A.2.5.1 of
the NSR-10 and herein in chapter 2.1.1.4)

12
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

(b) The design and construction of attached and vertical extensions, as well as other modifications of the
structural system
(c) The seismic rehabilitation of existing structures
(d) The design of seismic structural retrofit for structures damaged by an earthquake
It can also be applied for:
(e) Assessment of the seismic vulnerability of a building designed before the applicability of the current
Colombian code (NSR-10)
2.1.2.1.2 Step E2: Collection of information
Before starting the evaluation of the structure, the evaluating engineer must gather and study the available
information on geotechnical and structural design, construction process of the original structure and subse-
quent modification and must proceed to further investigations in accordance with chapter A.10.2.1 of the NSR-
10. The evaluating engineer must:
(a) Verify in-situ the validity of the information contained in the design documentation of the original
structure and its foundation system
(b) Investigate the structure in order to determine its state through evidence of local failures, cracks, ex-
cessive deflections, corrosion of rebars and other indicators of its behaviour
(c) Identify eventual settlements of the foundation and their effects on the structure
(d) Determine the potential occurrence in the past of extraordinary events that could have affected the
integrity of the structure (explosions, blazes, earthquakes, modifications of the structure or of building
allocation that increased the loads, and other modifications)
2.1.2.1.3 Step E3: Qualitative evaluation of the design and construction quality, and state of conservation
The evaluating engineer must provide a qualitative evaluation, for the structure and its foundation, of the orig-
inal design and construction, and of its state of conservation, according to chapter A.10.2.2 of the NSR-10.
He/she must provide a formal qualitative evaluation of:
(a) The quality of the design and construction of the original structure, with regard to the best technology
available at the time it was constructed
(b) The state of conservation of the structure, taking into account for the earthquakes that may have dam-
aged it, cracking due to imposed displacements, corrosion of the rebars, etc.
Both the quality of the design and construction, and the state of conservation must be qualified as (i) Good,
(ii) Regular, or (iii) Bad.
A major problem encountered by evaluating engineers in practice is the lack of information (documentation)
regarding the original design. Especially when dealing with structures built before the enforcement of the
CCC SR-84, it is difficult to assess issues as the best technology available at the time in the Colombian context.
Moreover, building documentation was not systematically collected at that time.
Finally, it seems illogical to assess the quality of the original construction according to the historical state-of-
the-art, since some phenomena (e.g. punching of slabs on columns) were greatly misunderstood, if not com-
pletely ignored, and could have a major influence on the rupture mode of the structure.
2.1.2.2 Vulnerability evaluation of the existing structure
The procedure presented here is described in chapters A.10.3 (“Design seismic movements with limited safety
level”) and A.10.4 (“Evaluation criteria for the existing structure”) of the NSR-10. The NSR-10 authorizes the
use of three alternative guidelines for the evaluation of the structure:
(a) “Seismic Evaluation of Existing Buildings”, ASCE/SEI 31-03, American Society of Civil Engineers,
Restron, Virginia, USA, 2003 [16]
(b) “Seismic evaluation and Retrofit of Concrete Buildings”, ATC-40, Vol. 1, Appendices, Vol2, Applied
Technology Council, Redwood City, California, USA, 1996 [17]
(c) “NEHRP Handbook for Seismic Evaluation of Existing Buildings”, FEMA 178, Federal Emergency
Management Agency / Building Seismic Safety Council, Washington D.C., USA, 1992 [18]

13
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

(d)
Figure 2.8 – Geographical zoning of the seismic hazard for limited safety level: Ae [2]
2.1.2.2.1 Step E4: Determination of the seismic hazard level and ground motions
For some cases of evaluation of existing structures (detailed below), the NSR-10 allows a limited safety level,
submitting the structure to lower design seismic loads. This loading corresponds to an earthquake with a prob-
ability of occurrence of 20% on a period of 50 years, i.e. a return period of 245 years. The factors Aa and Av
are replaced by a single acceleration factor Ae (see geographical zoning in Figure 2.8) in the formulation ex-
plained and illustrated in section 2.1.1.4 later in section 3.1.2.1.2. The author of the present report assumes
that this limited safety level has been introduced in order to allow the usage of buildings that could not comply
with the standard safety requirements. Indeed, as previously mentioned, an important share of the Colombian
building stock was built before the enforcement of the first Colombian design code, and such buildings are
quite unlikely to comply with the current safety level requirements, even with important rehabilitation. A re-
newal of this share of the building stock would generate unreasonable economical costs, and the limited safety
level allows to use this stock while maintaining some safety level.
The target safety level (limited or equivalent to one of a new structure), represented by the overstress and
flexibility indexes (explained in chapters 2.1.2.2.5 and 2.1.2.2.7), depends on the time the structure was orig-
inally designed and on its societal importance, as explained below:
Structures of Usage Groups III and IV:
(a) For edifications indispensable to the society and to community services (Usage Groups III and IV),
the target safety level is one equivalent to that of a new structure. They are hence subjected to an
equivalent seismic loading as described in chapters 2.1.1.4 and 3.1.2.1.2, and their overstress and flex-
ibility indexes shall not exceed 1.0. If the structure was already revised as a structure of Usage Group
III or IV under the validity of the NSR-98 (after the 19th February 1998), a revision according to the
NSR-10 is not obligatory.
Structures of Usage Groups I and II:
(b) For structures designed within the frame of the CCC SR-84:
i. If subjected to design seismic loading corresponding to that applied to new structures, a flex-
ibility index up to 1.5 is allowed, while the overstress index must remain under 1.0.
ii. If subjected to design seismic loading with limited safety level, both indexes of overstress and
flexibility must yield 1.0. This level of safety must be officially accepted by the owner.

14
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

iii. The rehabilitation interventions on non-structural elements can be limited to life-endangering


elements in case of earthquakes (e.g. façade risking to collapse on people, etc.).
(c) Structures designed before the validity of the CCC SR-84 must at least comply with the requirements
of the limited safety level, with both indexes of overstress and flexibility no larger than 1.0. This level
of safety must be officially accepted by the owner.
(d) For structures declared as historical heritage, provided they do not belong to Usage Group III or IV, a
safety level lower than the limited one (in the sense of chapter A.10.3 of the NSR-10 presented here)
is allowed. For this lower safety level to be accepted, the evaluating engineer must provide a justifi-
cation that severe restrictions prevent to reach a higher level of safety. It must be officially accepted
by the owner, and the reasons of this lowered safety level as well as the measures to restrict the public
access to the building and ensure the safety of its occupants must be stated in the construction license.
2.1.2.2.2 Step E5: Structural analysis
According to article A.10.4.2.5, the equivalent seismic loads Fs must be calculated through the (linear elastic)
equivalent lateral force method, following the procedure explained in chapter A.4 of the NSR-10 for new
structures. A dynamic analysis is allowed if the evaluating engineers deems the information on the existing
structure sufficient.
To do so, the structural system must be classified in the sense of chapter A.3 of the NSR-10 (herein chapter
2.1.1.5), in order to determine the energy dissipation factor. This energy dissipation factor for existing struc-
tures R’ must be defined by the evaluating engineer, with the following guidelines:
(a) When good information on the original design is available and sufficient, the evaluating engineer must
evaluate R’ by comparison with the guidelines of the NSR-10 for the energy dissipation factor for new
structures R in chapter A.3.
(b) When the information on the original design is not sufficient, the evaluating engineer must define a
value of R’ according to its best judgement. This value cannot be higher than the one established in
chapter A.3 for new structures having the same structural system and material.
(c) When there is no information, a value of ¾ of the one proposed in chapter A.3 for new structures of
the same structural system and the same material shall be taken for R’. The obtained value must not
be lower than 1.
(d) For non-reinforced masonry structures, the value of R’ must be taken as 1.
Regarding non-seismic loads, the standard recommendations for new structures must be applied, except that
dead loads must be estimated through on-site observations as well as according to the norm.
The combination of internal loads is done, as for new structures, according to eqs. (2.9) to (2.11), by simply
dividing the obtained internal seismic loads Fs by the energy dissipation factor R’, and combining these loads
with the other loads in load combinations LCULS_1 and LCULS_2, as for new structures. In comparison, the SIA
269 [19] uses load combinations with more favourable safety factors for the evaluation of existing structures
than for new structures, since the loads acting on an existing structures are generally better known than those
that will act on a future one (especially dead loads).
2.1.2.2.3 Step E6: Determination of the existing strength of all elements
The determination of the existing strength of the structure Nex for every element (for axial, shear, flexural and
torsional modes) must be determined according to the available information (in the pre-existing documentation
and further investigations carried out by the evaluating engineer) to his/her best judgement, according to its
expertise. The strength is defined as the ultimate strength of fragile materials, or the yielding point of ductile
materials. In general the strength considered in the evaluation of existing structures is defined through the same
process as the nominal strength of elements in new structures. It would however appear that no reduction factor
is used on the resistance.
2.1.2.2.4 Step E7: Determination of the effective strength of all elements
According to the evaluation of the original design and construction, and of the state of conservation (Good,
Regular, or Bad), two reduction factors are applied on the resistance (Table 2.3). ϕc corresponds to the reduc-
tion factor linked to the quality of the original design and construction, and ϕe represents the reduction due to
the state of conservation.

15
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

Table 2.3 – Values of ϕc and ϕe


Quality of the original design and
construction, or of the state of
Good Regular Bad
ϕ c or ϕ e 1.0 0.8 0.6
The effective (design) strength Nef is then obtained through eq. (2.15)
𝑁𝑒𝑓 = 𝜙𝑐 ∙ 𝜙𝑒 ∙ 𝑁𝑒𝑥 (2.15)
The reduction factors proposed here are global ones, acting on every element without rational distinction be-
tween the possible failure modes (e.g. shear vs. tension-controlled flexion vs. compression-controlled failure).
Those can even be equal to the unity, hence not guaranteeing any safety margin. This method seems to be quite
subjective and include a lot of uncertainties that lead to a large dispersion of the possible results. Recommen-
dations for possible improvements can be found in chapter 2.4.
2.1.2.2.5 Step E8: Determination of the overstress indexes
The overstress index ios of the whole structure is defined as the maximum overstress index ios,ij of all its struc-
tural elements in all type of loading (axial, shear, flexural, torsional):
𝑁𝑢,𝑖𝑗
𝑖𝑜𝑠,𝑖𝑗 = (2.16)
𝑁𝑒𝑓,𝑖𝑗
where Nu,ij and Nef,ij represent respectively the design load and strength of element i for loading type j.
It seems unfitting to consider a unique overstress index for the whole structure, moreover if this comes from a
linear elastic analysis, hence not taking into account for (generally favourable) load redistributions in the plas-
tic domain.
2.1.2.2.6 Step E9: Evaluation of the drifts
The drifts of the structure must be obtained from the analysis carried out in step E5 (chapter 2.1.2.2.2), using
the same procedure as for newly built structures (Step D9: 2.1.1.9).
2.1.2.2.7 Step E10: Determination of the flexibility indexes
Similarly to the overstress index, the flexibility index of the structure is defined as the maximum flexibility
index of all its floors. The flexibility index if,i of each story is obtained separately through eq. (2.17).
Δ𝑖
𝑖𝑓,𝑖 = (2.17)
Δ𝑎𝑑𝑚
where Δi is the inter-story drift of story i, and Δadm represents the drift limit (0.5 for fragile masonry construc-
tion, and 1% otherwise) imposed by the NSR-10 in chapter A.6 (herein chapter 2.1.1.10).
2.1.2.3 Intervention on the structural system
2.1.2.3.1 Step E11: Definition of the structural intervention
Once the vulnerability of the existing structure is determined by means of the overstress and flexibility indexes,
the nature of the structural intervention must be defined. Notably, if the overstress and flexibility indexes do
not comply with the requirements mentioned in chapter A.10.9 (herein 2.1.2.2.1), a repair/rehabilitation is
required. The different type of interventions must be classified as: (i) attached extension, (ii) extension in
elevation, (iii) actualization to the current regulation (and/or repair of damages), or (iv) modifications of the
structural system (other than extensions).
2.1.2.3.2 Step E12: Analysis of the final structure
Depending on the type of intervention, the chapter A.10 of the NSR-10 has different requirements regarding
the structural analysis of the final structure, including the designed intervention (extension, rehabilitation,
modification). For each category of intervention, the overstress and flexibility indexes must comply with the
requirements stated in Step E4 (chapter 2.1.2.2.1). If those requirements are not met, a rehabilitation/repair of
the existing structure must be carried out. In case of an extension, the new part of the structure must be sub-
jected to loads corresponding to the normal safety level for new structures. If the extension is vertical, both the
existing and the new structures must comply with the safety level corresponding to a new structure.

16
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

The evaluating engineer must provide a justification of all calculations and documents including:
(a) A list of all the design and construction documents of the original edification that were used during
the evaluation and the design of the modifications (architectural and structural plans, calculation notes,
soil studies, foundation design, inventory, construction diary, personal consultations with profession-
als that participated in the design or construction, etc.)
(b) A description of the evaluation of the state of conservation of the structure and of its foundation system
according to the requirements of A.10.2
(c) A clear description justifying the parameters used in the evaluation of the existing structure and the
design of the intervention
(d) Calculation notes of the proposed intervention with the corresponding justifications showing that the
final structure will achieve the requirements (strength and behaviour) of chapter A.10
(e) Other appropriate documents, as deemed appropriate by the evaluating engineer

2.2 Specific Requirements for Walls and their Seismic Design


2.2.1 General requirements for walls
2.2.1.1 Design method
The chapter C.14 of the NSR-10 offers three distinct design methods for walls from which the designer can
choose, with certain restrictions:
(a) Design the wall according to an empirical method based on a simple axial load check in chapter C.14.5.
This method can only be used for rectangular walls subjected to a compressive axial load with a small
eccentricity, such that the resultant stays in the inner third of the wall in both directions
(b) Design the wall as an element subjected to a combination of axial, flexural and shear (chapters C.10
and C.11). This is the most commonly used method. Section C.10.10 presents a method including
second-order effects by moment magnification to verify the out-of-plane stability.
(c) When flexural tension controls the out-of-plane design of a wall that has a regular section (no open-
ings), and is restrained against overturning at the top, an alternative method (C.14.8) based on the
iteration of deflections can be used to verify the stability instead of the analytical method presented in
chapter C.10.10. This method is presented as an alternative method for slender walls, sensitive to 2nd
order effects2.
2.2.1.2 Minimum thickness:
The NSR-10 (as the ACI) defines minimum walls thickness depending on the design method. For walls de-
signed as elements subjected to axial, flexural and shear walls according to chapter C.10 and C.11 (method (b)
above), no minimum thickness is defined. However, for normal walls designed with the empirical method
(C.14.5.3.1, method (a) above), a minimum thickness of (i) 1/25 of the wall length or height, whatever is
smaller, and (ii) no less than 100 mm is required, and 190 mm for exterior basement and foundation walls
(C.14.5.3.2).
For non-structural reinforced concrete walls (i.e. walls not considered in the structural analysis), the minimum
thickness is bounded by the 1/30 of the minimum distance between lateral supports, and 100 mm, whatever is
smaller (C.14.6.1). In practice, it seems design engineers extrapolate this minimum requirement to all RC walls
(including ones designed according to chapters C.10 and C.11), as no wall thinner than 100 mm was found
amongst the studied building in the frame of this report.
In comparison, the SIA 262 [13] imposes a minimum thickness of 150 mm for in site casted walls, and 100
mm for precast concrete walls (Table 21 in SIA 262 [13]), with a minimum thickness of 200 mm and 1/15 of
the story height for the boundary elements of shear walls resisting lateral seismic loading (article 5.7.1.2 of
262 [13]).
The EN 1992 sets a minimal thickness of 120 mm for in-situ casted walls (article 12.9.1 (1)). However, this
article is valid for plain and lightly reinforced concrete structures. It appears that no minimum thickness is
directly prescribed for normally reinforced concrete walls. For walls designed as ductile, the EN 1998-1 [20]

2
This is the interpretation of the author of the present paper. No indication is given in the NSR-10 to determine
if flexural tension controls the out-of-plane design of a wall, nor is there a definition of what is a “thin wall”.
17
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

states that the web must at least be 150 mm thick and 1/20 of the free story height (article 5.4.1.2.3), and article
5.4.3.4.2(10) sets a minimum thickness for the boundary element depending on its horizontal extension (be-
tween 1/15 and 1/10 of the free story height).
2.2.1.3 Maximum axial load ratio
When designing walls with the first method, there is no direct restriction in the NSR-10 on the maximum axial
load ratio. For elements designed with the empirical method (C.14.5, walls with low axial load eccentricity),
an indirect reduction of 45% of the axial resistance acts as an indirect restriction on the maximum axial load
ratio, as noticeable in eq. (2.18).
𝑘𝑙𝑐 2
𝜙𝑃𝑛 = 0.55𝜙𝑓𝑐′ 𝐴𝑐 [1 − ( ) ] (2.18)
32𝑡𝑤
where Pn is the nominal axial strength of cross section, ϕ stands for the strength reduction factor, fc’ is the
specified compressive strength of concrete, Ac denotes the gross area of concrete section, k is the effective
length factor, lc represents the length of the wall, measured between two lateral supports, and finally tw is the
wall thickness.
Finally, the only direct axial load ratio verification is defined for thin walls designed with the alternative iter-
ative method prescribed by chapter C.14.8 (method (c) above), where the NSR-10 prescribes a maximum load
ratio of 6% (C.14.8.2.6), but this limit is intended to act more as a restriction on the method validity than really
a way of limiting the axial load ratio of shear walls, since it is only applicable for the alternative iterative
method of C.14.8.
In comparison, article 5.7.1.1 of the SIA 262 sets a maximum axial load ratio of 40% (in design value) for
shear walls designed as ductile elements.
2.2.1.4 Reinforcement requirements
The reinforcement ratios for walls are prescribed in section C.14.3, provided that the following condition
(coming from C.11.9.8) is respected3:
𝑉𝑢 < 0.5𝜙𝑉𝑐 (2.19)
0.17𝜆√𝑓𝑐′ 𝑡𝑤 𝑑
or (free choice)
𝑉𝑐 = 𝑁
𝑁 𝑑 𝑙𝑤 (0.1𝜆√𝑓𝑐′ + 0.2 𝑢 ) (2.20)
𝑢 𝑙𝑤 𝑡𝑤
min (0.27𝜆√𝑓𝑐′ 𝑡𝑤 𝑑 + ; [0.05𝜆√𝑓𝑐′ + ] ℎ𝑑)
4𝑙𝑤 𝑀𝑢 𝑙𝑤
{ 𝑉𝑢 − 2
where Vu is the design shear load, Vc stands for the nominal shear strength provided by concrete (defined in
number C.11.2.1.2), Nu is the design axial load (taken positive in compression) combined with the design shear
load Vu and the design moment Mu, λ represents a modification factor reflecting the reduced mechanical prop-
erties of lightweight concrete, tw is the wall thickness, d stands for the structural depth of the element (here can
be assumed to be 0.8·lw, with the wall horizontal length lw).

3
After discussing with engineers members of the ACI committee in Colombia, it would appear that the factor
0.27 in the bottom condition of eq. (2.20) is a mistake, and should be replaced by 0.17, as in the upper condi-
tion.
18
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

The minimum longitudinal and transversal reinforcement ratios set by section C.14.3 depend on the rebar
diameter, and are respectively set according to eqs. (2.21) and (2.22). They are intended to prevent uncontrolled
spread of cracking and sudden failure, as well as a too big spacing of cracks, inducing larger crack openings.
 Longitudinally:
0.12% for bars Ø ≤ 16 mm
𝜌𝑙 ≥ { (2.21)
0.15% otherwise
 Transversally:
0.20% for bars Ø ≤ 16 mm
𝜌𝑡 ≥ { (2.22)
0.25% otherwise
where ρl is the longitudinal reinforcement ratio, ρt the transversal reinforcement ratio, and Ø the rebar diameter
in the considered direction.
If the requirement of eq. (2.19) is not fulfilled, the shear reinforcement must follow the requirements of section
C.11.9.9, which ensures that the reinforcement ratio guarantees structural safety when the shear capacity of
concrete is exceeded, and sets the minimum reinforcement ratios as follows:
 Transversally:
𝜌𝑡 ≥ 0.25% (2.23)
 Longitudinally:
𝜌𝑙 ≥ 0.0025 + 0.5(2.5 − ℎ𝑤 ⁄𝑙𝑤 )(𝜌𝑡 − 0.0025) ≥ 0.25% (2.24)
where hw is the total height of the wall and lw its length.
It is however not required for the longitudinal reinforcement ratio to be higher than the transversal one.
There is no upper limit for the reinforcement ratios prescribed by the NSR-10 for walls. For columns
(C.9.10.1), the reinforcement ratio must comply with eq. (2.25), which provides a maximum reinforcement
ratio, but there is no indication as to if the boundary element of a wall, its web or the wall as a whole should
be considered as a column. The definition of a column given in the code is that of an element with height to
minimum transverse dimension superior to 3, used principally to transfer axial compressive loads, which could
thus include the definition of a wall (and its boundary elements). However, that would be contradicted by the
imposed minimum reinforcement ratios of walls set by C.14.3 in eq. (2.23). A wall hence does not seem to be
considered as a column and is hence not subjected to this upper reinforcement ratio limit.
1% ≤ 𝜌 ≤ 4% (2.25)
Article C.14.3.4 sets an obligatory double-layer bidirectional reinforcement and distribution as well as pre-
scription of placement for walls thicker than 250 mm, to the exception of “basement walls”. It seems unclear
what the exact definition of basement walls is, as it is counter-intuitive to remove the requirement of double
layer of the reinforcement for the walls that are precisely the most solicited.
The maximum spacing for both longitudinal and transversal reinforcement is set in C.14.3.5 according to
eq. (2.26).
𝑠 ≤ min(3ℎ , 450 mm) (2.26)
where s is the centre-to-centre spacing of rebars, and h the wall thickness.
The section C.11.9.9, applicable when eq. (2.19) is not fulfilled, adds a supplementary limit for the maximum
spacing of lw/3 for the longitudinal reinforcement, and lw/5 for the transversal reinforcement.
Finally, article C.14.3.6 states that the longitudinal reinforcement does not need to be enclosed by stirrups
when the longitudinal reinforcement ratio does is not in excess of 1% or when the longitudinal reinforcement
is not required as compression reinforcement.

2.2.2 Specific requirements for seismic resistance of walls


In addition to the standard requirements applicable for every wall specified in chapter C.14, the NSR-10 de-
fines specific requirements for each structure type in chapter C.21, depending on its energy dissipation cate-

19
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

gory (DMI, DMO, DES) and structural system (frame, bearing walls, dual, etc.). It is allowed to design struc-
tural (and non-structural) members as not being part of the seismic load-bearing system (thus avoiding com-
pliance with chapter C.21), but their influence on the structural response and the consequences of their poten-
tial failure must be taken into account in the seismic design.
It is to be noted that the organization of chapter C.21 of the NSR-10 is slightly reorganized compared to the
corresponding chapter 21 in the ACI 318M-08. However, the content is vastly similar in both cases.
The present chapter describes briefly, in a non-exhaustive manner, the main requirements related to seismic
resistance for shear walls.
2.2.2.1 Materials in DMO and DES structures
In order to guarantee a sufficiently ductile behaviour, special requirements on the material properties of con-
crete and reinforcement steel (as well as mechanical and welded splices, and anchoring zones) are set by the
NSR-10 (sections C.21.1.4 to C.21.1.8) for DMO and DES structures. Amongst those requirements, the most
important are a minimum compressive strength fc’ of 21 MPa for normal concrete, a maximum compressive
strength fc’ of 35 MPa for light concrete, an actual yield stress fym of the steel not higher than 125 MPa above
the specified value fy, and a minimum fum/fym ratio of 1.25, where fum is the actual tensile strength of rebars.
2.2.2.2 DMI structural walls
According to C.21.1.1.7 b), structural walls of structures of category DMI do not need to comply with special
requirements from chapter C.21.
2.2.2.3 DMO structural walls
Most of the requirements for DMO walls (section C.21.4) are taken from the requirements for DES walls
(section C.21.9, called by article C.21.4.4), with a few changes explained in articles C.21.4.4.1 to C.21.4.4.4.
The present subsection will hence present the requirements for DMO walls and the following subsection
(2.2.2.4) will just introduce the additional requirements for DES walls when compared to DMO walls.
2.2.2.3.1 Connection requirements
When two walls or a wall and the foundation are connected in a DMO structure, section C.21.4.2 requires that
yielding of the connection is restricted to the steel parts (rebars or other) of the connection, maintaining the
concrete in the elastic regime. It is also required by C.21.4.3 that connecting elements or parts of elements that
are not designed to be ductile withstand a load of 1.5Sy, where Sy is the yield strength of the connection. These
specifications are set to guarantee the integrity of the connections when yielding occurs in adjacent regions.
Strangely, they seem to only be set for DMO walls, and not for DES walls, which do not seem to be subjected
to any requirements regarding wall connections.
2.2.2.3.2 Reinforcement requirements
According to C.21.9.2.1, the minimum reinforcement for the web is 0.25% in both directions, unless eq. (2.27)
is guaranteed, in which case the minimum requirements of chapter C.14.3 are sufficient.
𝑉𝑢 ≤ 0.083𝐴𝑐𝑣 𝜆√𝑓𝑐′ (2.27)
where Vu denotes the design shear load, and Acv is the gross area of the concrete section bounded by the web
thickness and the length of the section in the direction of the shear load considered.
The condition of eq. (2.27) corresponds (almost exactly) to that of eq. (2.19), provided that the estimation for
Vc is the upper condition in eq. (2.20). In other words, no real additional requirement is presented in this code
section, unless a refined estimation of Vc taking into account the axial load and moment (bottom condition of
eq. (2.20)) is considered in the general procedure for minimum reinforcement of walls. The present require-
ments are even less conservative in case hw/lw < 2.5 and the design shear load surpasses half of the shear
strength provided by concrete.
Article C.21.9.2.3 requires the use of minimum two layers of reinforcements when eq. (2.28), which compares
the shear design load with the shear strength provided by the concrete only, is not guaranteed.
𝑉𝑢 ≤ 0.17𝐴𝑐𝑣 𝜆√𝑓𝑐′ (2.28)
Indeed, if the design shear load surpasses the strength of the concrete section, shear cracks will spread, and
with only one layer of reinforcement and thin section, even the slightest eccentricity could lead to important
out-of-plane displacement, hence inducing potential buckling under axial/flexural as well as shear loading.

20
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

The use of two separate layers allows a better stability by providing a small lever arm to the tension induced
by shear loads.
In regions where yielding of the longitudinal rebars is expected, the anchoring length of rebars must be in-
creased by 25% in tension, according to C.21.9.2.4(c).
Finally, if the aspect ratio hw/lw of the wall is smaller than 2.0, the longitudinal reinforcement ratio ρl shall not
be lower than the transversal reinforcement ratio ρt, according to C.21.9.4.3, in order to resist shear loads in
both orthogonal directions (transversal and longitudinal).
2.2.2.3.3 Shear resistance
The shear load in a DMO structural wall shall not be higher than prescribed by C.21.9.4.1, illustrated in eq.
(2.29).

𝑉𝑢 ≤ 𝑉𝑛 = 𝐴𝑐𝑣 (𝛼𝑐 𝜆√𝑓𝑐′ + 𝜌𝑡 𝑓𝑦 ) (2.29)

where αc is a factor depending on the aspect ratio, equal to 0.25 for hw/lw ≤ 1.5, 0.17 for hw/lw ≥ 2.0, and varying
linearly in between.
It is to be noted that, whereas for frame columns of DES type (C.21.5.4.1) and beams of DMO type
(C.21.3.3(a)) the design shear load Ve is determined using the nominal (probable) moment resistance, hence
guaranteeing a flexural rupture, there is no such requirements for walls. Article C.21.9.3 states that the design
shear load Vu for walls must be calculated using the design load combination. This does not provide a guarantee
of flexural (and hence more ductile) rupture. However, chapter C.9.3.4(a) states that, for structures of type
DMO and DES, if the nominal shear resistance of the section is lower than the shear corresponding to the
development of the nominal flexural resistance, the resistance reduction factor ϕ should be lowered from 0.75
(C.9.3.2.3) to 0.6 when considering the shear resistance of the section under seismic loading. Even though this
procedure does not directly (nor completely) guarantee a flexural failure of the element through the application
of capacity design principles, it does lower the risk of a shear failure.
2.2.2.3.4 Axial and flexural resistance
Article C.21.9.5.1 states that walls subjected to a combination of axial and flexural loads must be designed
according to sections C.10, related to the design of elements subjected to flexion and axial load, with the
particularity that: (i) it is not necessary to take into account the nonlinear part of the stress-strain relationship
of concrete, (ii) the upper bounds given by eq. (2.30) for elements with spirals and eq. (2.31) for elements with
stirrups , which take into account the long-term effects need not be used, and (iii) the design flexural load Mu
need not be magnified in order to take into account instability (slenderness) effects.
𝜙𝑃𝑛(𝑚𝑎𝑥) = 0.80𝜙[0.85𝑓𝑐′ (𝐴𝑐 − 𝐴𝑠𝑡 ) + 𝑓𝑦 𝐴𝑠𝑡 ] (2.30)
𝜙𝑃𝑛(𝑚𝑎𝑥) = 0.75𝜙[0.85𝑓𝑐′ (𝐴𝑐 − 𝐴𝑠𝑡 ) + 𝑓𝑦 𝐴𝑠𝑡 ] (2.31)
where Ast is the total area of passive (non-prestressed) longitudinal reinforcement.
C.21.9.5.2 sets the maximum effective flange width of a wall section to 25% of hw, unless a more detailed
analysis shows a superior effective width.
2.2.2.3.5 Boundary elements requirements
According to C.21.9.6.1, C.21.9.6.2 (modified by C.21.4.4.1 for DMO walls) and C.21.9.6.3 (modified by
C.21.4.4.2 for DMO walls), design engineers have two different criteria presented below to assess the need of
using boundary elements, depending on the type of walls being designed (or assessed):
(a) For continuous walls with only one critical section for a combination of axial and flexural loads,
boundary elements are required when the highest neutral axis depth c under design horizontal dis-
placement δu is greater than the limit set by:
𝑙𝑤
𝑐≥ (2.32)
600(𝛿𝑢 ⁄ℎ𝑤 )
where the ratio δu/hw cannot be taken smaller than 0.0035. Note that δu should be computed according
to step D9 (herein section 2.1.1.9), but often an upper-bound is computed through the admissible inter-
story drift in the design process

21
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

When boundary elements are required according to eq. (2.32), they must be extended longitudinally
over a height of lw, but no less than Mu/4Vu.
or

(b) For discontinuous walls and alongside openings, boundary elements are required when the compres-
sive stresses generated by the load combinations including E (eqs. (2.10) and (2.11)) exceed 0.3fc’.
The boundary elements must be extended until the most compressed fibre reaches 0.22fc’4. These
stresses must be calculated with a linear elastic model of the structure using the gross section properties
subjected to design loads.
When boundary elements are required, C.21.9.6.4 (a) states that their horizontal length lo must satisfy
eq. (2.33).
𝑙𝑜 ≥ max(𝑐 − 0.1𝑙𝑤 , 𝑐/2) (2.33)
In sections containing flanges, the boundary elements must include the effective flange width and penetrate at
least 300 mm in the web, according to C.21.9.6.4 (a).
The maximum longitudinal spacing so of the confinement rebars shall not bet greater than set by eq. (2.34),
according to C.21.3.5.6(a) to (d) (called by C.21.4.4.3).
1
𝑠𝑜 ≤ min (8𝜙𝑙,𝑚𝑖𝑛 , 16𝜙𝑜 , max ( 𝑑𝑜 , 150 mm)) (2.34)
2
where ϕl,min is the minimum longitudinal rebar diameter in the boundary element, ϕo denotes the confinement
rebar diameter and do is the minimum dimension of the boundary element.
According to C.21.3.5.7 (called by C.21.4.4.3), the amount of confinement steel rebars must guarantee
eq. (2.35).
𝑠𝑜 𝑏𝑐 𝑓𝑐′
𝐴𝑠ℎ ≥ 0.06 (2.35)
𝑓𝑦𝑡
where Ash is the total area of the stirrups (on a perpendicular section of height so), bc denotes the transversal
dimension of the boundary element, measured from the external border of the confinement stirrups, and fyt
represents the specified yield stress of the stirrups.
According to C.21.3.5.7 (called by C.21.4.4.3), the minimum diameter of the confinement rebars is 10 mm,
and the maximum centre-to-centre spacing hx between two rows of confinement rebars (stirrups or hooks) shall
not be greater than 350 mm.
An inconsistency is present in article C.21.9.6.4(d), supposed to define the penetration requirements (in the
wall supports) of the transversal reinforcement of boundary elements. It refers to article C.21.9.2.3 as to the
required depth of penetration but said article speaks of something completely different (number of reinforce-
ment layers in function of the shear load).
Article C.21.9.6.4(e) (as presented in the errata published on March 19th, 2010 [8]), demands that the transverse
reinforcement of the web be extended such that it can develop its yield strength fy within the boundary element,
and must be extended at least until 150 mm from the extremity of the wall.
Finally, article C.7.1.4 states that when a boundary element is required in DMO and DES shear walls, the
confining stirrups must end in a minimum of 135° hook that penetrates at least 6ϕo into the section. Additional
hooks (used to maintain the distance between two opposite layers of reinforcement) are exempt from that
requirement. Those can have an angle of just 90°, even though those cross ties with an angle of only 90° are
known to open more easily and thus offer a reduced degree of confinement once the spalling of cover concrete
occurs, as explained by Paulay and Priestley’s book: Seismic Design of Reinforced Concrete and Masonry
Buildings from 1992 [21].

4
No indication is given as to whether this limit is set for the longitudinal or transversal extension of the bound-
ary element. The author supposes it refers to the longitudinal extension since the transversal extension is reg-
ulated by chapter C.21.9.6.4.
22
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

When no boundary elements are required, article C.21.9.6.5 states:


(a) If the longitudinal reinforcement ratio in the extremity of the wall is higher than 2.8/fy (with fy in
MPa), the transversal reinforcement in the extremity must comply with C.21.6.4.2 and
C.21.9.6.4(a). These articles set requirements for stirrups to prevent buckling of the longitudinal
bar.
(b) Except when Vu in the plane of the wall is lower than that prescribed in eq. (2.27), the transversal
reinforcement which ends in the extremities of structural walls without boundary elements must
have a standard hook that couples the border longitudinal rebar or the longitudinal rebar must be
coupled by a U-shaped pin of same spacing and diameter as the transverse reinforcement.
2.2.2.4 DES structural walls
The requirements for DES walls are exactly the same than those for DMO walls, except for the points presented
below.
2.2.2.4.1 Connection requirements
The requirements presented for connections between intersecting walls or between a wall and the foundation
presented in chapter C.21.4.2 and C.21.4.3 for DMO walls appear to not be required for DES walls, which is
apparently nonsensical.
2.2.2.4.2 Reinforcement requirements
No change in the requirements between DMO and DES walls.
2.2.2.4.3 Shear resistance
No change in the requirements between DMO and DES walls.
2.2.2.4.4 Axial and flexural resistance
No change in the requirements between DMO and DES walls.
2.2.2.3.5 Boundary elements requirements
In eq. (2.32) (C.21.9.6.2), the minimum value of the ratio δu/hw cannot be taken smaller than 0.007.
In the second criteria (C.21.9.6.3) for the need of boundary elements and their vertical extension, the corre-
sponding limits for are respectively 0.2fc’ and 0.15fc’.
According to C.21.6.4.3 (called by C.21.9.6.4(c)) the maximum spacing for confinement reinforcement must
comply with eq. (2.34).
1 350 − ℎ𝑥
𝑠𝑜 ≤ min ( 𝑑𝑜 , 6𝜙𝑙,𝑚𝑖𝑛 ,100 mm ≤ 100 + ≤ 150 mm) (2.36)
3 3
In DES structures, article C.21.9.9 stipulates that the reinforcement of columns that support discontinuous
walls must comply with the requirements for DES frame structures (article C.21.6.4.6).

2.3 Guidelines for the Pushover Analysis


Appendix A-3 of the NSR-10 offers guidance for engineers wishing to proceed with a pushover analysis of
the structure. The recommendations contained in that chapter are not compulsory and are meant to promote
discussion within the Colombian civil engineering community in order to implement properly the pushover
analysis in a future version of the code.
According to the recommendations of the NSR-10, the pushover method must take into account:
 The spatial distribution of mass and stiffness in the structure
 25% of the design live loads in the mass calculation (unlike normal design procedure)
 The nonlinearity of the elements once the proportionality displacement limit is exceeded
 The P-Δ effects
 The effect of axial load if it exceeds 15% of the member compression strength
 The strain-hardening, ductility, resistance and stiffness degradation of elements in the plastic domain
 The simulation of cracking in reinforced concrete and masonry elements

23
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

Figure 2.9 – Pushover analysis procedure schematic representation


When modelling the structure, the base must either be assumed fixed, or a more refined model taking into
account the stiffness and capacity of the foundations according to principles of soil mechanics must be imple-
mented.
In the presence of an attic, the control point for the pushover analysis is to be set at the centre of mass of the
story at the base of the attic. For buildings without attics, the control point is to be set at the centre of mass of
the highest level of the structure.
The loads are to be applied at the centre of mass of each story, distributed proportionally according to the
distribution from a modal analysis for the fundamental mode. At step j, the total lateral load Vj and the dis-
placement δj of the control node and the internal loads of all elements must be calculated and stored. The first
step of the pushover analysis must remain in the elastic range. The analysis must be carried out until at least a
displacement of 150% the target displacement δT (defined below).
A bilinear curve is to be adapted to the capacity curve, such that the first segment of the curve crosses the
capacity curve at 60% of the effective curve capacity at yielding δy (defined below), that the point correspond-
ing to the target displacement δT coincide for both curves, and that the area below the two curves between the
origin and the target displacement be equal. The effective yield strength Vy and the effective curve capacity at
yielding δy are defined by intersection of the two segments of the bi-linear curve.
The effective fundamental period Te and target displacement δT are defined through eqs. (2.37) to (2.41) pre-
sented below.

𝑉1 ⁄𝛿1
𝑇𝑒 = 𝑇1 √ (2.37)
𝑉𝑦 ⁄𝛿𝑦

𝑇𝑒 2 (2.38)
𝛿𝑇 = 𝐶𝑜 𝐶1 𝑆𝑎𝑒 ( ) 𝑔
2𝜋
𝑛
∑𝑖=1 𝑚𝑖 𝜙𝑖 (2.39)
𝐶𝑜 = 𝑛
∑𝑖=1 𝑚𝑖 𝜙𝑖2
1 (𝑅𝑑 − 1)𝑇𝐶 (2.40)
(1 + ) if 𝑇𝑒 < 𝑇𝐶
𝐶1 = {𝑅𝑑 𝑇𝑒
1.0 if 𝑇𝑒 ≥ 𝑇𝐶
𝑆𝑎𝑒 𝑀𝑔 (2.41)
𝑅𝑑 =
𝑉𝑦
where T1, V1 and δ1 are respectively the fundamental period of the structure, the base shear and control dis-
placement determined with the first step of the analysis, Co is a factor linking the equivalent Single Degree Of
Freedom system (SDOF) displacement with the fundamental vibration mode of the Multiple Degree Of Free-
dom system (MDOF), C1 represents a factor taking into account the inelastic behaviour in the system response,
and Sae depicts the value of the elastic design acceleration spectrum for the effective period Te. Sae is expressed
in multiple of the earth gravity acceleration g. mi and ϕi denote respectively the mass (including 25% of the
design live loads) and shape function of the fundamental vibration mode for story i, Rd stands for the ductility
24
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

factor of the system, TC symbolizes the vibration period corresponding to the transition between the constant
and decreasing acceleration segments of the spectrum, and finally M is the total building mass, including 25%
of the design live loads.
This procedure is largely based on the Displacement Coefficient Method (DCM) as implemented in the
Prestandard and Commentary for the Seismic Rehabilitation of Buildings, FEMA-356 [22]. However, the
FEMA-356 takes into account two additional modification coefficients (C2 and C3) in the determination of the
target displacement. The coefficient C2 takes into account the degradation of stiffness and strength degradation
over the hysteretic response, depending on the framing type of the structure, the period of the structure and the
target performance level (Immediate Occupancy, Life Safety or Collapse Prevention). The coefficient C3 rep-
resents the increased displacements due to dynamic P-Δ effects. No justifications are given in the NSR-10 for
disregarding those two coefficients. It is also to be noted that the NSR-10, unlike the FEMA-356, does not
explicitly state that the effective yield shear strength of the building Vy shall not be taken greater than the
maximum base shear force over the complete curve. Another difference can be found in the calculation of the
ductility factor in eq. (2.41). In the FEMA-356, the effective mass is considered through a factor Cm, whereas
the NSR-10 procedure promotes the use of the complete building seismic mass.
Some of the recommendations regarding the analysis also come from the Recommended Seismic Design Cri-
teria for New Steel Moment-Frame Buildings, FEMA-350 [23], notably regarding the loads that need to be
taken into account, and the drift level until which the analysis must be carried out.
According to article A-3.2.9, if eq. (2.42), comparing the yield strength to the majored base shear load, is
respected, it is not necessary to proceed to a complete analysis of the whole structure.
𝑉𝑦 > Ω𝑜 𝑉𝑠 (2.42)
𝑉𝑠 = 𝑆𝑎𝑒 𝑔𝑀 (2.43)
were Ωo is the overstrength factor given in tables A.3-1 to A.3-4 of the NSR-10 depending on the structural
system, and Vs symbolizes the shear load at the building base, with Sae calculated for the effective period Te.
If the condition of eq. (2.42) is not guaranteed, the engineer must carry a complete analysis of the whole
structure, evaluating the internal loads and deformations of all elements and connections of the structure for
the load step corresponding to the target displacement. The capacity of elements and connections must be
evaluated with the help of data from laboratory tests on similar components. The effect of gravity loads on the
deformation capacity must be taken into account. Moreover, the deformation of an element supporting gravity
loads cannot be greater than (i) 2/3 of the deformation corresponding to the loss of vertical bearing capacity,
and (ii) 2/3 of the deformation corresponding to a loss of 30% of the peak strength5.
In addition to the abovementioned requirements, the structure must be designed such that no reduction of the
global lateral load Vj occurs before a displacement of at least 125% of the target displacement δT is reached.
Finally, an independent committee composed of at least 2 certified structural engineers must revise the seismic
design.

2.4 Final Comments on the NSR-10


The first comment to be made about the NSR-10 (and the ACI 318M-08 by the same occasion) is that the
amount of descriptions of different analysis procedures and possible design method that it contains makes it
sometimes closer to a textbook than an actual standard, supposed to set the minimum requirements and main
guidelines to design a building. While a certain level of restrictions and instructions are necessary and desira-
ble, especially if the expertise of engineers varies greatly within a population, an overly directive and detailed
code tends to reduce the liberty of innovation of the designer and make its application more cumbersome.
Furthermore, through the different methods and design criteria, numerous cross references, exceptions, and
sometimes contradictions in the code make reading and understanding of the code more difficult, hence po-
tentially causing errors due to misinterpretations of the code.
Besides, as mentioned in the introduction of the present chapter, the structural concrete chapter of the NSR-10
is largely based on the ACI-318-08, with only a few minor changes. This can be a source of potential problems,

5
It is safe to assume that the peak strength evoked here is the lateral one.
25
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

especially since construction practices in Colombia are very different than those in the U.S. This is especially
true regarding the quality of the material and its regularity, the type of elements used, as well as the in-situ
conditions of construction and conservation of the different elements and parts of elements (for instance when
it comes to precast piers drilling, still sometimes effectuated manually in Colombia).
Some additional comments on specific aspects of the NSR-10 should also be made.
Apart from the minimum thickness associated to the use of some design methods, there is no prescription of
minimum thickness for walls. This is especially surprising for walls of DMO and DES structures, since a
thinner wall will be more sensitive to the biaxial loading resulting from seismic loading, often not considered
in design. A thin wall will also be sensitive to instability effects, which can have significant influence on the
rupture mode and reduce drastically the displacement capacity. Additionally, the magnification of moments
due to out-of-plane behaviour is disregarded in seismic design (C.21.9.5.1), hence increasing the risk of such
issue to occur in practice. For this reason, setting a minimum thickness for all walls, regardless of the design
method, would be advisable, eventually setting a different (higher) minimum thickness for boundary elements
of walls designed as ductile elements, in the spirit of the SIA 262 [13] (minimum thickness of 150 mm for the
web of walls according to articles 5.5.4.1 and 5.7.1.3, and 200 mm for the boundary element according to
article 5.7.1.2).
Similarly, there is no direct restriction for shear walls regarding the maximum axial load ratio they can be
subjected to, nor maximum reinforcement ratio, in spite of the reduction of ductility they may induce, through
early crushing of the concrete under relatively low lateral displacements, not to mention the potential instabil-
ities caused by high axial load ratios. Few experimental evidence is however available to assess the exact
implications of both high axial load ratios and high reinforcement ratios in regard to seismic performance, and
further research would therefore be needed to evaluate appropriately the consequences of such combination.
Meanwhile, setting a maximum axial load ratio of 40% like the SIA 262 is advisable, at least for structures of
category DES.
While the SIA recommends avoiding lap splices in plastic regions, the NSR-10 provides similar guidance only
for frame elements of categories DMO and DES, and does not require such directives for walls, despite the
experimental and in-situ evidences of the poor behaviour of lap splices in heavily cracked regions, as explained
by Paulay and Priestley 1992 [21]. A simple recommendation in that sense could promote a safer design, while
not imposing too strict constructive complications.
While it is required in DMO structures to design the connections between intersecting walls, as well as between
walls and foundation so that the yielding localizes in the steel components, this condition is oddly disregarded
for DES structures, and would be easy and straight forward to implement. This is highly recommendable, since
intersecting walls, while often designed separately, tend to work together and their intersections are often
highly solicited under bi-axial loading as explained by Paulay and Priestley 1992 [21].
For DMO and DES structures, the NSR-10 requires to consider the moment resistance to determine the design
shear load for beams and columns. However, it is not demanded for the design shear load in walls to be related
to the development of the maximal moment. It is difficult to understand the logic behind this difference, espe-
cially since the moment resistance is proportional to the structural depth to the square, while the shear re-
sistance increases proportionally to the structural depth. Walls, having a larger structural depth, would hence
tend to be more subjected to brittle shear failure than columns (all other things being equal).
The use of single-layer reinforcement is largely spread in Colombia, sanctioned by the NSR-10, which only
requires two layers of reinforcement when the shear load is within a specified lower limit, or when the wall
thickness is high. This is of concern in view of the proven unsatisfactory behaviour associated to out-of-plane
instability shown by Paulay and Priestley’s article from 1993: Stability of Ductile Structural Walls [24].
The flat-rate limits imposed for inter-story drifts in the seismic analysis are not related to an actual rational
limit (e.g. damage to the non-structural elements, crushing of concrete, start of heavy cracking, instability
effects). While convenient for a quick verification, it seems unclear what the motivation for these precise
admissible drift limits is. This is especially true for damage control earthquakes prescribed for structures of
Usage Groups III and IV; these are meant to guarantee that the structure does not enter the plastic domain
under light seismic loading, and the simple verification of flat-rate admissible drifts does not seem to ensure
it. Granting the design/evaluating engineer the freedom to set other limits, based on the capacity of structural
and non-structural elements, could be advised.

26
Colombian RC wall buildings assessment – Review of the Colombian design code for buildings

While allowing the engineer to use more advanced methods (e.g. pushover, linear dynamic analysis), the ap-
proach proposed by the NSR-10 is mainly based, both for the design and evaluation of structures, on a simple
equivalent lateral force method (i.e. a force-based, linear elastic approach). This method, although beneficial
for its simplicity and relatively straightforward process, does not provide the engineer with a proper apprecia-
tion of the seismic behaviour of a structure in case of larger seismic events than expected (i.e. larger than the
design earthquake). Notably, it does not ensure a ductile behaviour in such cases, which is especially concern-
ing for the evaluation of existing structures which comply with the limited safety level, with a shorter return
period for the design earthquake. A capacity-based method would be advisable, especially in the evaluation of
existing structures, where the engineers need to assess the actual capacity of a building in order to determine
the necessity and nature of a structural intervention.
In a similar spirit, it seems unfitting and overly conservative to use global indexes for the evaluation of the
whole structure, especially when based on an elastic approach. Indeed, this does not take into account the load
redistribution in the plastic domain, and does not provide information as to the real ductility demand vs. ca-
pacity of each elements. Indeed, an element with an overstress index of 1 can either be close to collapse or far
from it, depending on its state of deformation, ductility, and strain-hardening properties. The abovementioned
implementation of a nonlinear analysis of the structure, considered as a whole, would provide a better under-
standing of the need for reinforcement and rehabilitation of a structure. This could take into account the exist-
ing capital of resistance and more important displacement capacity, considering plastic redistributions, and
could induce important rehabilitation cost reductions, while providing a more accurate representation of the
structural capacity.
Finally, the reduction factors proposed for the strength evaluation of existing structures are flat-rate ones, not
related to a failure mode (e.g. shear, compression, traction), and dangerously ignore the potential of fragile
rupture of elements, especially if the qualitative evaluation of the original design, construction and state of
conservation are considered as good by the evaluating engineer. Moreover, the evaluation of the design and
construction quality is supposed to be based on the state-of-the-art at the time of the construction, hence ig-
noring scientific findings that might have occurred afterwards on potentially dangerous failure mode (e.g.
punching). This is particularly concerning for structures built before the introduction of the first Colombian
code, as the definition of state-of-the art in Colombia at the time of the construction becomes highly arbitrary.
The author would defend a more rational approach, with a reduction factor depending on the quality of the
available information (including investigations made by the engineer), a reduction factor depending on the
failure mode in the same spirit as for the design procedure, based on the current state-of-the-art, and finally a
coefficient for the state of conservation, varying depending on the observations and investigations carried out
by the evaluating engineer.

27
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

3 SELECTION AND DESCRIPTION OF THE REFERENCE BUILDING

The present chapter describes the preliminary analysis carried out during the procedure for the selection of the
reference building, provides a presentation of the selected structure, and describes the results of its preliminary
analysis.
In order to better appreciate the Colombian construction practices, a preliminary gathering of information has
been carried out by Prof. Blandón and Prof. Bonett with a team of students. Information was collected on a
number of buildings located in the municipalities of Medellín and Envigado, in the department of Antioquia,
Colombia (region of intermediate seismic hazard). Amongst those buildings and after some quick hand calcu-
lations, it was decided to run simple preliminary analyses on a set of 6 buildings of 9 to 17 stories height, in
order to select the most fitting building to represent Colombian design practices.

3.1 Selection of the Reference Building


3.1.1 Desired outcomes of the detailed analysis and selection criteria
The objectives of the detailed numerical analysis are numerous and complex, they must hence be defined
clearly for a proper characterization of the choice criteria in the reference building selection.
3.1.1.1 Desired outcomes of the detailed analysis
The general desired outcome of the analysis is an evaluation of the behaviour of typical Colombian buildings.
This sizeable task requires the analysis of more than one building. In this sense, the first objective of the
detailed numerical analysis presented in this report is to provide an assessment of the validity of simplified
analysis methods in the Colombian context, with the hope that those could be later carried out for a large set
of buildings, hence allowing a broader evaluation of Colombian design practices. Verifications of the simpli-
fied analysis methods to calculate the fundamental period T, displacement demand, yield displacements and
ductility demand are of special interest.
In the scope of the present thesis, the following questions are of further interest:
 Estimation of the building behaviour in case of a design earthquake and in case of an earthquake of
higher order of magnitude
 Evaluation of the “real” drift of the building under design seismic loads
Note that the assessment of the building behaviour in case of an earthquake of higher order of magnitude is of
special interest because of two reasons:
 The design requirements of the NSR-10 for RC walls (presented in chapter 2.2) vary very little be-
tween DMO and DES structures, and the design of a building located in a region of intermediate
seismic hazard (where DMO is the minimum requirement) could be very similar in essence to this of
a building located in a region of high seismic hazard (where DES is required).
 Given the variability of seismic events, the design earthquakes only have a limited degree of reliability,
and testing the robustness of a building design is always recommendable.
3.1.1.2 Selection criteria
Having the abovementioned objectives in mind, the first selection criterion for the reference building is of
course the availability and quality of the data (plans, material information, address, etc.).
Secondly, it is highly important that the building presents a regular profile with standard Colombian wall
configuration, for if its characteristic are too unusual, the results of the present analysis would be too specific
and only represent the behaviour of the studied building, hence prohibiting extrapolations to the general Co-
lombian building stock.
Finally, in the spirit of continuity between the pre-study and the present thesis, buildings with walls presenting
features comparable to the ones tested at EPFL (low thickness, one layer of reinforcement, presence of walls
intersections, heavier reinforcement at the extremities of the walls) are preferable.

28
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

Figure 3.1 – Representation of the preliminary analysis procedure

3.1.2 Preliminary analysis procedure


In order to obtain a range of values for the fundamental period of the buildings, as well as their displacement
and ductility demand, different approximations and hypotheses were made regarding both the fundamental
period estimation procedure (I, II and III) and the soil type corresponding to each building (soil type C or D).
Figure 3.1 shows in a diagram the procedure to calculate the outputs of the preliminary analyses with the
different hypotheses.
3.1.2.1 Evaluation of the displacement demand
In order to evaluate the displacement demand, the fundamental period of the building must first be assessed,
to identify the building equivalent single degree of freedom system (SDOF) in the prescribed response spectra
of the code.
3.1.2.1.1 Estimation of the building fundamental period
Three levels of estimations were used to assess the fundamental period of the building: I, II and III, with each
level supposed to provide a closer estimation to the "real" fundamental period. The three estimations followed
the indications of chapter A.4.2 of the NSR-10.
The first empirical estimation for the fundamental period TI shown in eq. (3.1) presents the simplest formula,
based solely on the amount of stories N of the building.
𝑇I = 0.1 ∙ 𝑁 (3.1)
where N is the number of stories in the building.
The second estimation shown in eq. (3.2) is slightly more advanced, as it takes into account for the total height
of the building as well as the structural system type (here: reinforced concrete shear walls).
𝑇II = 𝐶𝑡 ∙ 𝐻𝑛𝛼 (3.2)
where TII is the second estimation of the building fundamental period, Ct and α are factors depending on the
structural system (respectively 0.049 and 0.75 for RC shear walls), and Hn is the total building height.
Finally, the most refined estimation uses a similar expression, but uses an empirical formula to take into ac-
count the floor area of building stories, as well as the web area and length of walls resisting in the considered

29
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

direction, as shown in eqs. (3.3) and (3.4). This estimation is especially useful since it takes (partly) into ac-
count the actual structural system, hence allowing to differentiate the period estimation for perpendicular di-
rections in plan.
𝑛𝑤
100 𝐻𝑛 2 𝐴𝑤𝑖
𝐶𝑤 = ∙∑ ( ) (3.3)
𝐴𝐵 ℎ𝑤𝑖 ℎ 2
𝑖=1 1 + 0.83 ( 𝑤𝑖 )
[ 𝑙𝑤𝑖 ]
0.0062 (3.4)
𝑇III = ∙ 𝐻𝑛1.00
√𝐶𝑤
where TIII is the third estimation of the building fundamental period, AB is the story area of the building, nw
represents the number of walls resisting in the considered direction, hwi is the total height of wall i, lwi stands
for the horizontal length of wall i in the considered direction, and Awi is the shear area of the section of wall i,
in the direction considered (i.e. the area of its web).
3.1.2.1.2 Estimation of the design displacement demand according to the NSR-10
In order to estimate the demand, the spectrum proposed by the NSR-10 has been adopted. Acceleration and
velocity coefficients of Aa = 0.15 and Av = 0.20, corresponding to the urban area of Medellín, have been con-
sidered, according to table A.2.3-2 of the NSR-10. Those coefficients can also be found in Figure 2.1 of the
present report.
As a first assumption, soils of categories C and D have been taken into account, resulting in amplification
coefficients of respectively Fa = 1.2 and 1.5, and Fv = 1.6 and 2.0, according to tables A.2.4-3 and A.2.4-4 of
the NSR-10.
All buildings studied belong to Usage Group I, resulting in a coefficient of importance of I = 1.00.
The characteristic periods of the spectrum can then be calculated according to eqs. (3.5) to (3.7), with To only
being useful in case of a dynamic analysis for superior vibration modes, hence not used here.
𝐴𝑣 ∙ 𝐹𝑣
(𝑇𝑜 = 0.10 ∙ ) (3.5)
𝐴𝑎 ∙ 𝐹𝑎
𝐴𝑣 ∙ 𝐹𝑣 (3.6)
𝑇𝐶 = 0.48 ∙
𝐴𝑎 ∙ 𝐹𝑎
𝑇𝐿 = 2.4 ∙ 𝐹𝑣 (3.7)
where To, TC and TL are the characteristic periods of the demand spectrum.
The elastic displacement demand spectrum Sde is constructed according to eq. (3.8), as illustrated in Figure
3.2. The elastic response has been used for the preliminary analyses in a spirit of simplicity. The use of the
elastic displacement response instead of the actual plastic demand corresponds to the equal displacement rule
largely accepted throughout the engineering community.
0.62 ∙ 𝐴𝑎 ∙ 𝐹𝑎 ∙ 𝐼 ∙ 𝑇 2 if 𝑇 < 𝑇𝐶
𝑆𝑑𝑒 = { 0.3 ∙ 𝐴𝑣 ∙ 𝐹𝑣 ∙ 𝐼 ∙ 𝑇 if 𝑇𝐶 < 𝑇 < 𝑇𝐿 (3.8)
0.3 ∙ 𝐴𝑣 ∙ 𝐹𝑣 ∙ 𝐼 ∙ 𝑇𝐿 if 𝑇𝐿 < 𝑇

30
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

Figure 3.2 – Elastic displacement demand spectrum [2]


3.1.2.2 Evaluation of the displacement ductility demand
The displacement ductility demand has been evaluated for the main seismic-resistant walls of each building.
It is largely based on the work of Priestley et al. (Displacement-Based Seismic Design of Structures [25]),
more precisely on chapters 4.4.7 and 6.2.1, with some adaptations for geometrical configurations that were not
exactly covered by this reference.
3.1.2.2.1 Estimation of the yield curvature and displacement
In chapter 4.4.7 of [25], the estimation of the yield curvature of typical reinforced concrete wall shapes pre-
sented in eq. (3.9) is proposed.
𝜀𝑦 2.00 for rectangular walls
𝜙𝑦 = 𝐶𝑖 ⁄𝑙 with 𝐶𝑖 = { (3.9)
𝑤 1.50 for flanged walls
where ϕy is the wall yield curvature and εy denotes the yield strain of steel rebars, typically εy = fy/Es =
420/200’000 = 2.1‰, with the steel rebars Young’s Modulus Es = 200’000 MPa.
From thise equations, it was estimated that a factor of Ci = 2.00 could be used for double-T and C walls with
relatively symmetric flanges, loaded in the direction of the web. It was also considered that Ci = 1.50 could be
used for L-shaped walls and C-shaped walls loaded in the direction of the flanges. Finally, for C and double-
T walls with notably asymmetric flanges, loaded in the direction of the web, eq. (3.10) has been adapted from
eq. (3.9).
𝜀
𝜙𝑦 = 𝐶𝑖 𝑦⁄𝑙 with 𝐶𝑖 = 1.75 for assymetric C and double-T walls (3.10)
𝑤
In order to verify the validity of the yield curvature estimation, a sectional analysis software called Re-
sponse2000 has been used to assess the moment-curvature curves of each important wall under an axial loading
ratio of 5%, roughly corresponding to typical solicitations at the base of building RC walls.
According to chapter 6.2.1 of [25], the effective height He of the SDOF used to model the building can be
taken, in a first analysis, as 70% of the total building height Hn. With this assumption of a SDOF of height He
loaded by a single concentrated load at its top, the yield top displacement of the SDOF can be calculated
through the unit force theorem. Given this SDOF has a linear curvature profile, the yield displacement Δy is
determined through eq. (3.11)
𝐻𝑒
1
̅ ∙ 𝑑𝑠 =
Δ𝑦 = ∫ 𝜙ℎ ∙ 𝑀 ∙ 𝜙 ∙ 𝐻2 (3.11)
3 𝑦 𝑒
0
̅ denotes the moment profile over the SDOF
where ϕh is the curvature profile over the SDOF height, and 𝑀
height, due to the unit force.

31
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

3.1.2.2.2 Estimation of the plastic displacement and ductility demand


The plastic displacement demand Δp was simply calculated according to eq. (3.12), using the equal displace-
ment rule. This rule is generally used for the response of systems with periods larger than the period corre-
sponding to the end of the constant acceleration plateau TC, but the extension of this hypothesis for shorter
periods is sufficient for the precision required in the preliminary analyses.
Δ𝑝 = 𝑆𝑑𝑒 − Δ𝑦 ≥ 0 (3.12)
Finally, the displacement ductility demand μΔ for the SDOF was calculated through eq. (3.13).
𝑆𝑑𝑒
𝜇Δ = ⁄Δ (3.13)
𝑦

This procedure was applied for each building, for each major wall in each direction, for the three estimations
(I, II and III) of the building fundamental period calculation, and with the assumption of a soil class C and D.
For the sake of clarity and reading facility, only the results of the preliminary analysis for the selected building
are presented.

3.1.3 Final comments on the reference building selection


Amongst the 6 buildings studied in the preliminary analyses, sufficient information to run a detailed analysis
of the building was present for 5 of them. According to the preliminary analyses presented in section 3.1.2, at
least one wall entered the plastic domain according to the estimation level III for the period and the hypothesis
of a soil of quality D. The estimated displacement ductility demand was however low in all buildings (μΔ <
1.65 for all of them). The selected reference building was chosen for the clarity and quality of the information
in the available plans, for the regularity of its profile, and for the diversity of its walls (lengths and geometric
configurations). Finally, it was also selected because its walls presented similar features to the wall tested at
the EESD laboratory: low thickness, single layer of reinforcement and no confining stirrups.

3.2 Reference Building Description and Preliminary Analysis Results


3.2.1 Denomination convention
In order to facilitate the writing and reading of the present report, the following conventions will be used:
 The intersection of two axes (e.g. axis G and 2) will be denoted G|2
 The direction corresponding to axes denoted by a letter will be referred to as direction X and the
direction corresponding to axes denoted by a number will be called direction Y.
 The wall numbering can be observed in Figure 3.3. This numbering corresponds to the numbering
used by the design engineer.

3.2.2 Description of the reference building


The selected reference building, located in the hills of the municipality of Medellín, consists of a 9-stories
housing building whose structural system is composed of shear walls of different lengths (0.75 to 5.7 m) and
shapes (Rectangular, L, T, C, double-T), all of which have a thickness of tw = 100 mm and are reinforced by a
single bidirectional welded mesh placed at the centre of the wall. All the meshes are of type D131, i.e. com-
posed of rebars of diameter Ø5 mm placed at a distance of 15 mm in both directions. The reinforcement ratio
in both direction is hence for all walls ρ = 0.131%. Figure 3.3 shows the building’s floor plan as well as the
wall nomenclature. The walls filled in with grey were selected for the preliminary analysis in section 3.2.3
and/or the drift limit definition in section 5.3.1. All floors have a height of 2.30 m, and all slabs are 0.10 m
thick. At the base of the walls, foundation beams of dimension 0.30 x 0.50 m2 are placed. The foundation is
made of abovementioned beams and piles.
Concrete quality varies between the different stories, with a concrete of strength f’c = 35 MPa for the first floor,
f’c = 28 MPa for the second floor and f’c = 21 MPa for the superior levels as well as for other structural elements
(slabs, beams, piles etc.). The steel meshes have a yield strength of fy = 490 MPa, and the additional rebars (in
slabs, piles and beams) reach fy = 422 MPa. The concrete cover is of 7.50 cm in foundations, 2.50 cm in beams
and 2.00 cm in slabs.

32
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

Figure 3.3 – Floor plan


33
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

Figure 3.4 – Elevation of wall 1


34
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

The building presents two irregularities. The first is a discontinuity in direction Y at axis L in the depth of the
building base due to the sloping ground, as illustrated in Figure 3.4. The second irregularity, of far lesser
importance, is a small difference in length between two opposing walls in regard to the symmetry axis L. Wall
I (which will herein be referred as H), located at Q|2, is 0.3 m longer than wall G (1.65 m for I vs 1.35 m for
G), located at G|2.
Since the building is located in a region of intermediate seismic hazard level, the energy dissipation capacity
category of the building is set as DMO. Its Usage Group is I, as stated above. A uniform live load of 1.8 kN/m2
for apartments and parking lots and a non-structural dead load of 1 kN/m2 for the finishings are admitted by
the designer. In addition, table B.3.4.3-1 of the NSR-10 recommends a minimum uniform dead load of 3.0
kN/m2 for buildings having masonry non-structural walls, and a live load of 0.5 kN/m2 is prescribed in table
B.4.2.1-2. This, combined with the self-weight of the 100 mm thick slab, results in a total dead load of 6.5
kN/m2 for the apartment floors and 3.5 kN/m2 for the roof.
The soil on which the building is located is composed of colluvial settlements made of mud flows and debris.
A microzoning campaign [26] has revealed the soil properties for seismic events shown in Table 3.1, which
produces the (elastic) response spectrum of Figure 3.5 through eqs. (3.14) and (3.15). Figure 3.5 also indicates
the periods estimated by the preliminary analysis and the spectra recommended by the norm that were used
during the preliminary analysis which are shown in Table 3.2.
Table 3.1 –Spectral coefficients for damage control and design earthquakes [26]
a s,max S ae,max /I To TC Fa α
Earthaquake Spectrum
g g s s - -
Damage control 0.05 0.23 0.10 0.50 4.50 1.43
Design 0.27 0.70 0.10 0.60 2.60 1.34
𝑇
(1 + (𝐹 − 1)) 𝑎𝑠,𝑚𝑎𝑥 ∙ 𝐼 for 𝑇 < 𝑇𝑜
𝑇𝑜 𝑎
𝑆𝑎𝑒 = 𝐹𝑎 ∙ 𝑎𝑠,𝑚𝑎𝑥 ∙ 𝐼 for 𝑇𝑜 ≤ 𝑇 ≤ 𝑇𝐶 (3.14)
𝛼
𝑇𝐶
{ ( ) 𝐹𝑎 ∙ 𝑎𝑠,𝑚𝑎𝑥 ∙ 𝐼 for 𝑇𝐶 < 𝑇
𝑇
𝑆𝑎𝑒 𝑇2
= 𝑆 𝑎𝑒 ∙ 𝑆𝑑𝑒 = (3.15)
𝜔2 4𝜋 2
The soil properties coming from the microzoning campaign presented in Table 3.1 correspond, for the funda-
mental period estimation II and III in direction Y, to a soil of lesser quality than category C and D used in the
preliminary analysis. However, for the fundamental period estimation I and III in X, the spectrum coming from
the microzoning campaign indicates a displacement demand similar to the demand for a soil of category C.

Elastic response spectrum


0,5
Elastic displacement demand [m]

0,4
Soil type C
0,3 Soil type D
Microzoning 2011
0,2 T
I
T II
0,1 T III,X
T III,Y
0
0 0,5 1 1,5 2 2,5 3 3,5 4
Period [s]

Figure 3.5 – Elastic displacement response spectra


35
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

3.2.3 Preliminary analysis of the reference building


The period estimation, as well as the design displacement coming from the preliminary analysis can be found
in Table 3.2 and are represented in Figure 3.5, while the plastic displacement and displacement ductility de-
mand for the major (i.e. largest) walls in each directions are presented in Table 3.3. Walls selected for the
preliminary analysis are filled in with grey in Figure 3.3. The factor Ci mentioned in Table 3.3 is the one used
in the calculation of the yield curvature, in eqs. (3.9) and (3.10). Both Table 3.2 and Table 3.3 also indicate
the displacement demand coming from the microzoning campaign with the indication Micro.
According to the supposedly more accurate estimations (III) for the period, the walls in direction X enter
slightly in the inelastic regime (ductility demand of 1.29 and 1.23 for soil type D), while walls in direction Y
are still by far in the elastic regime (50% of the yield displacement). This can be explained by the density of
walls, more important in the Y direction, hence stiffening severely the building in that direction. Two problems
having opposite effects are to be noted concerning the preliminary analysis:
(a) The (best) estimation of the fundamental period of the building were made here with formulas that do
not take flanges of walls into account, and the major walls in direction X all have important flanges,
which should increase significantly the stiffness of the building, hence reducing the demand in this
direction
(b) The soil quality is lower than the one used in the preliminary analysis, this underestimated the dis-
placement demand in both direction.
Table 3.2 – Period and displacement demand estimations for the reference building
Period Displacement demand
s m
S de, I, C 0.086
TI 0.90 S de, I, D 0.108
S de, I, Micro 0.082
S de, II, C 0.027
T II 0.49 S de, II, D 0.034
S de, II, Micro 0.042
S de, III, X, C 0.088
T III, X 0.91 S de, III, X, D 0.109
S de, III,X, Micro 0.083
S de,III,Y,C 0.023
T III, Y 0.45 S de, III, Y, D 0.028
S de, III,Y, Micro 0.035

36
Colombian RC wall buildings assessment – Choice and Presentation of the Reference Building

Table 3.3 –Displacement ductility demand evaluation for the major walls of the reference building
Wall Type Direction lw Ci ϕy He Δy S de Δp μΔ Hypothesis
- - - m - 1/m m m m m - Period Soil
0.086 0.002 1.02 C
0.108 0.023 1.27 I D
0.082 0.000 0.96 Micro
0.027 0.000 0.32 C
H
A X 3.3 1.75 0.0011 15.12 0.085 0.034 0.000 0.40 II D
assymetric
0.042 0.000 0.49 Micro
0.088 0.003 1.03 C
0.109 0.025 1.29 III D
0.083 0.000 0.97 Micro
0.086 0.000 0.97 C
0.108 0.019 1.21 I D
0.082 0.000 0.92 Micro
0.027 0.000 0.30 C
B T X 2.7 1.5 0.0012 15.12 0.089 0.034 0.000 0.38 II D
0.042 0.000 0.47 Micro
0.088 0.000 0.98 C
0.109 0.021 1.23 III D
0.083 0.000 0.93 Micro
0.086 0.030 1.54 C
0.108 0.052 1.92 I D
0.082 0.026 1.46 Micro
0.027 0.000 0.48 C
1 Rectangular Y 5.7 2 0.0007 15.12 0.056 0.034 0.000 0.60 II D
0.042 0.000 0.75 Micro
0.023 0.000 0.40 C
0.028 0.000 0.50 III D
0.035 0.000 0.63 Micro

3.3 Summary of the Reference Building Choice


The selected building, composed of thin walls reinforced in their centre by a single reinforcement layer, has 9
stories and appears, according to the preliminary analysis, to enter only slightly in the plastic zone (in one
direction) when subjected to design earthquake recommended by the NSR-10.
Most reinforced concrete constructions are able to sustain relatively important plastic deformations. However,
the construction practices in Colombia, well represented by this building, often present lightly reinforced con-
crete walls, while having a very low thickness. This, through the lack of confinement and enhanced sensitivity
to instability effects, implies a low ductility of the wall (when compared with other construction practices),
failing under low plastic deformation. This is notably shown in chapter 4 of the present thesis, presenting a
wall tested at the EESD laboratory of the EPFL in December 2013 presenting similar features to the walls
present in the reference building. This is of concern since the design requirements for RC walls in regions with
high seismic hazard (DES) are not significantly different from the requirements the reference building was
subjected to (DMO). The demand spectra however can change drastically, with a spectrum that can double in
intensity between Medellín and Vilavicencio for instance. These significant variations of demand, combined
with design practice lacking of robustness lead to increased risks of important damage and eventually building
collapse, especially since seismic events are in their nature very variable in intensity.

37
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

4 NUMERICAL MODELLING OF SINGLE WALL: VALIDATION WITH THE EX-


PERIMENTAL RESULTS

The aim of this chapter is to validate the numerical modelling approaches used to simulate the reference build-
ing behaviour. To this end, two models were used to predict the response of a wall tested in December 2013
in the EESD laboratory at the EPFL: (i) a nonlinear shell model using the software VecTor2 (V2), based on
the Modified Compression Field Theory (MCFT) by Vecchio et al. [27] and the Disturbed Stress Field Model
(DSFM) by Vecchio [28], and (ii) a nonlinear beam-element model using the software SeismoStruct (SS) [29].
The wall was tested in a combination of shear and flexural loads, with a constant axial load ratio. The experi-
mental and numerical results were compared at the global and local levels to assess the validity of the analyses.

4.1 Experimental Tests on a Thin RC Wall Representative of Colombian


Construction Practice
The wall Thin Wall 1 (TW1) was tested under cyclic loading. However, for the sake of time saving in postpro-
cessing of results, the Vector2 and SeismoStruct models were only subjected to monotonic loading in each
direction. Indeed, it is safe to assume that the monotonic response of the wall is the backbone of the cyclic
one. The tested wall is a 4/5 scale representation of a typical wall for the new emerging design and construction
trend in Colombia.

4.1.1 Geometry and reinforcement


The wall, as illustrated in Figure 4.1 and Figure 4.2, has a monosymmetric T-shaped section. It has a length lw
= 2700 mm and height hw = 2000 mm. Its web and flange have a thickness tw = 80 mm. The flange has a width
bf = 440 mm. An upper and a lower beam of thickness t = 440 mm simulate the effects of concrete slabs, and
provide supports to load the specimen. Both beams are heavily reinforced (in comparison to the wall) and
subjected to an external prestress in order to minimize the propagation of cracks in them, prevent uplift and
ensure that concrete compression stays within reasonable limits. The flange is present to simulate the effect of
another perpendicular wall on lateral stability, concrete confinement, as well as load transferring mode (e.g.
concentration of damage). During the experiment as well as during the numerical modelling, the horizontal
displacement of the control point indicated in Figure 4.2 has been tracked. It is located above the centroid of
the wall section, at the height of the actuators applying the horizontal displacement, i.e. at mid-height of the
upper beam.

Figure 4.1 – 3D rendering of TW1

38
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Figure 4.2 – TW1 geometry (including position of the centroid), dimensions in [mm]
The wall itself has a base reinforcement composed of a single central mesh of rebars of diameter Ø6 (mm)
placed at a distance dh = 240 mm of each other horizontally and dv = 200 mm vertically. Additionally, rebars
of diameter Ø6 are placed horizontally in the plane of the flange at the height of the web horizontal reinforce-
ment. In the flange, 4 extra Ø6 are placed symmetrically (2 on each side of the symmetry axis of the section).
There is a lap splice of 350 mm of the vertical Ø6 at the bottom of the wall.
In addition to this base reinforcement, 3 Ø16 where placed at a distance dh = 100 mm of each other on each
side of the wall these bars do not present any lap splices within the wall itself.
The reinforcement layout is illustrated in Figure 4.3 and the reinforcement ratios are shown in Table 4.1. Note
that the term "boundary element" refers to the zone of the web were the extra Ø16 are placed (300 mm on each
side of the wall).
Table 4.1 - Reinforcement ratios of TW1
Longitudinal Transversal Orthogonal
Zone
ρ l Ø6 ρl Ø16 ρ l Total ρ t Ø6 ρ o Ø6
Foundation beam - - - - -
Top beam - - - - -
Web 0.147% - 0.147% 0.177% -
Flange 0.321% 0.571% 0.892% 0.177% 0.177%
Boundary element - flange 0.161% 2.285% 2.445% 0.177% -
Boundary element - web 0.118% 2.513% 2.631% 0.177% -

39
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Figure 4.3 – Reinforcement layout

4.1.2 Material properties (MP)


Three distinct materials where used in the wall: the concrete, the steel of rebars Ø6, and the steel of rebars
Ø16.
4.1.2.1 Concrete
Four cylinders, produced with the concrete cast in the wall, have been tested in compression at the Laboratoire
de Matériaux de Construction (LMC) at the EPFL. The tests were produced according to the standards of the
Swiss Society of Engineers and Architects (SIA). The test results can be found in Table 4.2.
Table 4.2 - Uniaxial co mpression test results

Cylinder
Volumetric mass hs ds Ultimate load f c' Ec
t/m3 mm mm kN MPa MPa
2054 2.36 315 16 544 27.2 N/A
2055 2.35 315 16 596 29.8 25000
2056 2.37 315 16 570 28.5 25500
2057 2.37 315 16 560 28 25500
Average 2.36 315 16 568 28.4 25333

40
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Table 4.3 – Double Punch test results

Cylinder
hs ds ft'
mm mm MPa
1 153 80 2.53
2 153 80 2.05
3 153 80 1.82
4 153 80 2.39
Average 153 80 2.2
In order to determine the tensile strength of concrete, a double punch test has been carried out by the EESD
laboratory on 4 cylinders of diameter ds = 80 mm and height hs = 153 mm, with a diameter of punch dp = 20
mm. The results of said test can be found in Table 4.3.
4.1.2.2 Steel rebars
Two different steel were used for rebars of diameter Ø6 and Ø16. 8 Ø6 rebars have been tested in tension in
the EESD laboratory at the EPFL. The average properties resulting from the tests are presented in Table 4.4
and the detailed results of the tests are shown in Figure 4.4. Striction appeared within the measuring zone only
on tests number 2 and 4 for the Ø6 rebars, and on test number 1 for the Ø16 rebars. It is noteworthy that Ø6
rebars showed no yielding plateau.
Table 4.4 - Average material properties of tested Ø6 and Ø16 rebars
fy εy ε sh fu εu Es
Rebars
MPa mm/m mm/m MPa mm/m MPa
Ø6 460 2.3 2.5 625 80 183500
Ø16 565 2.825 27 650 110 208100

Tensile test Ø6 and Ø16 rebars


700
Ø6 TEST 1

600 Ø6 TEST 2
Ø6 TEST 3
500 Ø6 TEST 4
Ø6 TEST 5
Stress [MPa]

400
Ø6 TEST 6
Ø6 TEST 7
300
Ø6 TEST 8
200 Ø16 TEST 1
Ø16 TEST 2
100 Ø16 TEST 3
Ø16 TEST 4
0
0 50 100 150 200 250 Ø16 TEST 5
Strain [mm/m] Ø16 TEST 6

Figure 4.4 – Stress-strain curves of the tensile test on the Ø6 and Ø16 rebars

41
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Figure 4.5 – Schematic representation of the loading process

Figure 4.6 – Loading direction convention

4.1.3 Loading
The tested wall represents typical walls in many buildings in Colombia that have a shear span of typically
Ls = M / V = 10 to 18 m, and an axial loading ratio of approximately N/Acfc’ = |σc|/fc’ = 5 to 10 %. Following a
serie of numerical and hand calculation of the wall response under different loadings, it was decided to subject
the wall to an axial load ratio of 5 %, and apply a combination of lateral force and top bending moment in
order to obtain a shear span of 10 m, representing a shear span of 12.5 m in full scale (the tested wall scale is
of 4/5), see Figure 4.5.
The wall was subjected to a quasi-static cyclic loading with increasing lateral displacement with two loading
cycles per drift levels. The chosen loading direction convention can be seen in Figure 4.6. Positive loading
direction corresponds to a displacement opposed to the flange, and negative loading direction corresponds to
a displacement towards the flange.

4.2 Modelling of the Wall with Two Distinct Finite Element Approaches
VecTor2, presented in section 4.2.1 is a very refined software able to simulate the majority of the reinforced
concrete behavioural concrete. Its usage is very time-consuming (for the preprocessing, computing time and
the postprocessing). Moreover, it cannot model three-dimensional structures. This makes it a very useful tool
to verify the response of separate elements while being unable to model a complete structure. SeismoStruct,
presented in section 4.2.2, on the other hand is well-adapted to model three-dimensional complete structures.
However, the strong hypotheses on which it is based (e.g. Euler-Bernoulli beam theory, constant confinement
factor over each element, etc.) reduces the validity of its prediction. Both softwares were used in parallel in

42
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

the comparison with the experimental data in order to: (i) verify the validity of the numerical modelling options
used in the SeismoStruct model of the complete reference building presented in chapter 5.1, and (ii) to assess
the value of certain key input parameters for this model (e.g. confinement coefficient).

4.2.1 VecTor2 Model


VecTor2 is a nonlinear finite element program based on the Modified Compression Field Theory, used to
model reinforced concrete membranes. It is very able to model most of the concrete behavioural features.
However, due to its complexity, it is quite time-consuming to model the structure and postprocess the results.
4.2.1.1 Model geometry
Since, for the VecTor2 model, the height of the model per se is not as much of a limiting factor as for the
experimental test, the wall could be modelled with its full 10m in VecTor2, in order to be able to apply directly
a displacement at the top guaranteeing a shear span of Ls = 10 m.
After several modelling attempts, it was observed that the use of concrete elements with smeared reinforce-
ments rather than discrete reinforcing bars did not affect much the results, and even gave a more stable model.
This simplified greatly the postprocessing of the results, reason why smeared reinforcements have been chosen
in this analysis, separating each type of reinforcement (direction and diameter) in a different reinforcement
layer for each material. The lap splice of the Ø6 rebars at the base of the wall was however not modelled,
which is considered not too significant.
The model’s geometry and material numbering convention is visible in Figure 4.7. All elements used are plane
stress rectangular elements. All elements are rectangular displacement-based type elements, with width of 80
mm for the flange and 100 to 120 mm (depending on the reinforcement pattern) for the rest of the wall, and a
height of 100 mm. All material types shown in Figure 4.7 correspond to a different reinforcement configuration
and/or thickness of the section. For instance material type 3 correspond to the wall web, with its base rein-
forcement and its thickness of 80 mm, while material types 5 and 6 have the same thickness but contain the
additional Ø16 rebars, they hence have a higher vertical reinforcement ratio, and material type 4 has an en-
hanced thickness of 440 mm.
Material 2 10 m (Loading height) Δloaded

Node no 2653
Material 4

Material 5

Material 3

Material 6

Node no 702

2.2 m (Horizontal actuator height) Δtracked

0 m (Base of the wall)


Material 1

Figure 4.7 – VecTor2 model geometry and material numbering

43
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

4.2.1.2 Material properties


All material types are concrete reinforced with 0 to 4 smeared reinforcement layers in order to model each
type of rebars layers in each direction, respecting each time their respective reinforcement ratios. For the top
and foundation beams, it was assumed sufficient to increase the tensile and compressive strength of concrete
to a high value in order to keep cracks from propagating, without modelling their reinforcement and/or pre-
stress (here, fc’ = ft’ = 99 N/mm2). Table 4.5 summarizes all (relevant) selected material models used in the
VecTor2 model.
Following subsections (4.2.1.2.1 and 4.2.1.2.2) present the important material properties defined in the model
for both Concrete and Rebars.
Table 4.5 - List of material behaviour model used for the VecTor2 analysis
Concrete Steel Rebars
Phenomenon Model Phenomenon Model
Compression Pre-Peak Hognestad (Parabola) Stress-Strain Response Non-linear strain-hardening
Compression Post-Peak Modified Park-Kent Buckling Refined Dhakal-Mackawa
Compression Softening Vecchio 1992-A (e1/e2-Form) Hysteretic Response Bauschinger Effect (Seckin)
Tension Stiffening Modified Bentz 2003 Dowel Action Tassios (Crack slip)
Tension Softening Linear
Confined Strength Kupfert / Richart
Dilation Variable - Kupfer
Cracking Criterion Mohr-Coulomb (Stress)
Crack Stress Calculation Basic (DSFM/MCFT)
Crack Width Check Agg/2.5 Max crack width
Crack Slip Calculation Walraven (Monotonic)
Creep and Relaxation None
Hysteretic Response Nonlinear w/ Plastic Offsets

4.2.1.2.1 Concrete
After testing several specimens as discussed in 4.1.2.1, the material properties of concrete have been set ac-
cording to Table 4.6.
Table 4.6 – Concrete material properties for the VecTor2 model
f c' ft' Ec εo ν sx sy
MPa MPa MPa mm/m - mm mm
perpendicular reinforcement
28.4 2.2 25'333 2 0.2
spacing

4.2.1.2.2 Steel rebars


After the tests presented in 4.1.2.2 were carried out, the steel’s properties for both rebar types were set as
presented in Table 4.7. Since the experimentally tested wall was subjected to cyclic loading whereas the Vec-
Tor2 model was only subjected to monotonic loading, it was decided to neglect the yield plateau of the Ø16
rebars, as this plateau disappears once the rebars enter their plastic domain and are subsequently cycled. It has
also been assumed that the elastic Young’s modulus of the steel was exactly Es = 200’000 [N/mm2].
The material properties for steel rebars, as well as the properties if the yield plateau had not been neglected
can be observed in Table 4.7 and Figure 4.8. The figure also displays the experimental results of some of the
tests carried out on the rebars, as presented in section 4.1.2.2.
Table 4.7 - Steel rebar material properties for the VecTor2 model (with and without yield plateau)
Es εy fy ε sh εu fu
Rebar
MPa mm/m MPa mm/m mm/m MPa
Ø6 200'000 2.3 460 2.51 80 625
Ø16 200'000 2.83 565 2.83 110 650
Ø16 - With yield plateau 200'000 2.83 565 27 110 650

44
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Stress-strain relationship of rebars for the Vector2 model


700

600

500
Stress [MPa]

400 Ø6 TEST 4
Ø16 TEST 1
300 Ø6
Ø16
200
Ø16 - With yield plateau

100

0
0 50 100 150
Strain [mm/m]

Figure 4.8 – Stress-strain curves for steel rebars for VecTor2 model and selected experimental results
4.2.1.3 Load application procedure
As previously mentioned, the VecTor2 model was subjected to monotonic horizontal loading (imposed lateral
displacement) in both directions (+ and -), combined with the application of an axial load ratio of 5 %.
In order to ensure that the axial load is applied on the centroid of the wall, each node at the top of the wall was
subjected to a vertical load. This vertical load was set separately for each top node of the model to generate a
uniform stress of |σc|= 5/100·fc’ over the whole section. The rebars were not considered to influence in an
important way the position of the centroid, and were hence not taken into account in this distribution.
The horizontal loading was carried out by applying a horizontal displacement at the node no 2653 in the (ge-
ometric) centre of the wall, at a height of 10 m. This displacement Δloaded (see Figure 4.7), starting at 0 mm at
the beginning of the loading, was increased of 1 mm at each load step, until a total displacement of 400 mm
was reached.
In order to compare with the experimental results, the horizontal displacement Δtracekd (see Figure 4.7) of node
no 703 was tracked. This node is located at the same X-coordinate as node no 2653 but at a height of 2.2 m,
corresponding to the height of the horizontal actuator in the experiment.

4.2.2 SeismoStruct Model


SeismoStruct (SS) is a structural finite element software for dynamic analysis of frame structures. It is able to
predict the large displacement of structures under quasi-static loadings, as well as to compute their vibration
periods or proceed to time-history analysis. However, since it is based on Euler-Bernoulli beam theory, it does
not explicitely take into account inelastic shear deformations.
4.2.2.1 Model geometry
The geometry of the SeismoStruct model is shown in Figure 4.9. It simply consists of one element of 2.2 m
height (in order to reach the height of the actuator). An inelastic frame force-based element (infrmFB) with 6
integration points (IP) was used. Its section (T-shaped, discretized in 301 fibers) corresponded to the shape of
the tested element, with all (vertical) rebars considered separately and placed at their actual location. The
transverse reinforcement can however not be modelled in SeismoStruct. Only the influence of hoops on the
stress-strain law can be indirectly accounted for through empirical confinement models. The software consid-
ers separately the cover and core concrete. Since this separation makes little sense in a configuration with only
one central layer of reinforcement, this distinction was not assumed in this modelling, and the 25 mm of
“cover” and the “core” concrete were assigned the same material. Note that the differentiation makes little (or
no) sense in the configuration of TW1 but is automatically embedded in the software.

45
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Figure 4.9 – SeismoStruct Model geometry and loads


4.2.2.2 Material properties
4.2.2.2.1 Concrete
The behaviour model chosen for concrete, presented in Figure 4.10, is Mander et al 1998 [30], with the refine-
ments of Martinez-Rueda and Elnashai 1997 [31] . According to the tests presented in 4.1.2.1, the strength of
concrete has been set to fc’ = 28.4 MPa. The strain at (unconfined) peak stress has been chosen as εo = 2 ‰.
The tensile strength of concrete has been disregarded (set to ft’ = 0 MPa) to take into account for the heavy
cracking occurring under cyclic loading of concrete, as well as to avoid convergence problems in the numerical
analysis.
Finally, a parametric analysis was carried out with the help of the VecTor2 model described in section 4.2.1 to
define the confinement factor. In this analysis, the stress-strain curves of concrete in the four bottom elements
at the free end of the web (shown in Figure 4.11) were plotted when pushing in the positive direction. The
stress-strain curves can be seen in Figure 4.12. The confinement factor for each element has then been deter-
mined according to eq. (4.1), coming from Martinez-Rueda and Elnashai 1997 [31].

Figure 4.10 – Stress-strain relationship of the concrete in the SeismoStruct model

46
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Flange side Free end of the web

4th element
3rd element
2nd element
1st element

Figure 4.11 – Elements selected for the parametric analysis to define the confinement factor
𝑓𝑐𝑐
𝑘𝑐 = (4.1)
𝑓𝑐′
where kc is the confinement factor, fcc is the confined concrete peak stress, and fc’ is the concrete cylinder
compressive strength.
This resulted to confinement factors of 1.32, 1.10, 1.05, and 1.05 respectively for the 1st, 2nd, 3rd and 4th element
from the bottom of the wall (each element has a height of 100 mm). Knowing that the damage in the experiment
was concentrated not directly at the base but a few centimetres above it, it has been decided to use a confine-
ment factor of 1.15 in the SeismoStruct model. This is an intermediate value between the confinement factor
of the 1st and 2nd element, hence representing approximatively the first 200 mm of the wall.

Figure 4.12 – Stress-strain evolution for the concrete of the bottom 4 elements of the free extremity of the
web in the VecTor2 analysis when pushing in the positive direction

47
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Figure 4.13 – Stress-strain relationship of the rebars in the SeismoStruct model (a) Ø6 (b) Ø16
4.2.2.2.2 Steel rebars
For both Ø6 and Ø16, the model Menegotto-Pinto 1973 [32], with the adjustments of Filippou 1983 [33] has
been chosen. The yield strength was set to fy = 460 and 565 MPa for respectively Ø6 and Ø16. The Young’s
modulus was left to its default value (Es = 200 GPa), as well as the other material parameters. The stress-strain
relationship of rebars can be seen in Figure 4.13.
4.2.2.3 Load application procedure
The static pushover (i.e. monotonic) analysis has been carried with a Response control procedure (i.e. with
each load step corresponding to a target drift level, while in Load control procedure each load step is corre-
spond to a target load level). The permanent axial load of N = - 347.6 kN (corresponding to an axial load ratio
of 5 %) has been combined with an incremental lateral load and top bending moment coordinated to ensure a
shear span of Ls =10 m at the bottom of the wall. The target displacement was set to 0.03 m in both directions
(+ and -), with 201 load steps. The capacity of the wall was only reached in the positive direction (pushing
away from the flange).

4.3 Comparison between Experimental and Numerical Response


The experimental and numerical response of TW1 were compared at the global (section 4.3.1) and local (sec-
tion 4.3.2) levels. Since the wall TW1 presents features similar to the walls of the reference building, this
comparison is used to assess the range of validity of the SeismoStruct analysis and adapt some key parameters
of the reference building model (e.g. the confinement factor). Later on, the results of the experimental test on
wall TW1 were used to define the failure criterion for the reference building.

4.3.1 Global level response


The global response, presented in Figure 4.14, shows the plot of the displacement (at the height of the actuator
Δtracekd: 2.2 m, see Figure 4.7) in function of the base shear for TW1. As it is to be expected, the experimental
failure occurred when pushing in the positive direction (opposite to the flange), by crushing of the concrete
under important out-of-plane displacement. Both the VecTor2 and the SeismoStruct analysis provide a good
approximation of the general envelope of the wall behaviour. However, since both analysis were monotonic
and not cyclic, they did not simulate the loss of stiffness due to the progress of cracking and residual strains
coming from previous cycles, as it can be observed in the displacement range of approximately 1 to 7 mm in
both directions.
It is important to notice that neither of the numerical analysis reproduced the rupture point of the wall (dis-
placement capacity) directly on the global response. This is notably due to the fact that out-of-plane move-
ments, that played an important role in the experimental test, are not considered in either of those analysis, as
well as because the degradation of the resistance due to cyclic loading was not taken into account. It is noted
that the SeismoStruct analysis predicted a much closer estimation of the wall displacement capacity (in the
positive direction) than the VecTor2 model (failure at respectively 24.5 and 75 mm versus 16 mm for the
experiment). However, the capacity predicted by SeismoStruct on a global level is highly dependent on the
selected confinement factor, which is difficult to estimate accurately (21.5 mm for a confinement factor of

48
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

1.04, 24.5 mm for 1.05, and 28.5 mm for 1.06, hence a variation of approximately 15% for the response re-
sulting from a difference of 1% on the input).

Global response comparison


200

150

100
Base shear [kN]

50

0 Experiment
VecTor2
-50
SeismoStruct
-100

-150

-200
-25 -20 -15 -10 -5 0 5 10 15 20 25
Horizontal displacement at the actuator height [mm]

Figure 4.14 – Base shear vs top horizontal displacement: Comparison between numerical models (VecTor2
and SeismoStruct) and experimental results

4.3.2 Local level response


For the local level response comparison, the strain evolution at the bottom corners of the wall (flange and web
side) have been plotted against the lateral displacement for the experiment, the VecTor2 analysis, and the
SeismoStruct model. The experimental data come from actuators that were placed at the extremities of the wall
measuring the average strain on the first 50 mm of the wall height. The VecTor2 strains are the strains of the
elements located on the wall corners, they represent hence average strains on the first 100 mm of height in the
element. The SeismoStruct strains are the uniaxial strains at the most extreme section fibres, at the bottom
integration section in the element. Figure Figure 4.15 and Figure 4.16 show the evolution of the strains post-
processed to plot only the strains coming from lateral loading. The strains have been normalized to remove the
uniform compressive strains coming from the axial loading, meaning that a strain of zero on the plots actually
corresponds to a strain of 5/100·fc’/Ec ≈ 5/100·28.4/30000 ≈ 0.05 mm/m, which is negligible.

49
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Strain Comparison - Flange


25

20
Vertical strain [mm/m]

15

10 Experiment
VecTor2
5 SeismoStruct

-5
-25 -20 -15 -10 -5 0 5 10 15 20
Horizontal displacement at the actuator height [mm]

Figure 4.15 – Vertical stress at the base vs top horizontal displacement comparison, flange side

Strain Comparison - Web


25

20
Vertical strain [mm/m]

15

10 Experiment
VecTor2

5 SeismoStruct

-5

Horizontal displacement at the actuator height [mm]

Figure 4.16 – Vertical stress at the base vs top horizontal displacement comparison, web side
The first remark to be made is that the strain predictions are closer to the measured ones on the flange side.
This is due to two separate effects somehow related: (i) on the flange side, less damage appeared, the cracks
were more regular and distributed, and there was no (perceptible) out-of-plane movements, hence ensuring a
more regular and representative recording of the experimental data, given the measurement method, and (ii)
since there was no out-of-plane movements on the flange side, the hypothesis on which the numerical analysis
are based (e.g. Bernoulli’s beam theory: plane sections remain plane, etc.) were more representative of reality.
The first justification is also the most likely reason for the wavy experimental pattern results apparent in Figure
4.16, that are hence more attributable to the measurement method than to a real strain evolution.

50
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Secondly, it is noticeable that the strains are generally better represented in the VecTor2 analysis than with
SeismoStruct. This is easily explainable by the fact that SeismoStruct is a program developed primarily for
frame structures (i.e. with beam elements), and buildings built with shear walls can only be partly simulated
with this approach. It is hence logical to have some significant difference for the local response, especially in
the plastic domain. VecTor2, on the other hand, with its refined mesh over the two dimensions of the wall and
evolved 3D material model is able to represent more accurately the strain distribution. However, SeismoStruct
seems to be able to simulate relatively well the compressive strains on the flange side. This is however not the
type of locations where the critical local deformations are to be expected in the building, since the failure is
more likely expectable on free ends of walls, and is hence not of significant use.

4.4 Conclusions on the Validity of the Numerical Modelling Approaches


It comes out of this chapter that the global response seems to be simulated with sufficient accuracy both in the
SeismoStruct and the VecTor2 analyses. However, neither of the two models were able to simulate directly the
failure point of the wall and hence its capacity with a sufficient reliability. It is to be noted that wall TW1
reached failure after significant (noticeable with the bare eye) out-of-plane movements were observed, while
none of the two used software are able to simulate such instability effects that can lead to early failure of
elements.
The local level response simulation was, both for VecTor2 and SeismoStruct, of lesser accuracy than the global
response. However, while VecTor2 showed a simulation with sufficient level of accuracy to provide relatively
reliable failure criteria (i.e. concrete strains limits), the predictions of the SeismoStruct model were not satis-
factory enough to define such failure criteria.
In conclusion, SeismoStruct seems to be able to simulate well the envelope of the global response of a structure
built with walls similar to the ones included in the selected reference building. However, it does not provide
sufficient accuracy in the local response simulation to define strain limits as failure criteria. It is hence recom-
mended to carry a global analysis of the building with SeismoStruct to determine the shear distribution amongst
the walls and define what the displacement demand on each wall is, and then define a suitable drift limit. Part
3 of the Eurocode 8 (EN 1998-3 [34]) presents an empirical formula in its appendix A (refined in the 2013
corrigendum [35]) to define the chord rotations for different limit states. Combined with experimental tests
with similar features as the studied building, these limits can be used to define suitable drift limits. It has been
considered that wall TW1 presented enough similarities with the walls of the reference building to define drift
limits, as discussed in section 5.3.1.

51
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

5 NUMERICAL MODELLING OF REFERENCE BUILDING: DESCRIPTION,


ANALYSIS AND RESULTS

Figure 5.1 – Reference Building SeismoSrtuct model


This chapter presents the model used to analyse the building and the outcomes of the numerical analysis,
notably its capacity to withstand the design displacement demand, and its behaviour in case of an earthquake
of higher importance than prescribed by the NSR-10.

5.1 Description of the Reference Building Model


The Reference building has been analysed with a nonlinear beam-elements model using the software Seismo-
Struct (SS). The details of the model, displayed in Figure 5.1, are presented in the following subsections.

5.1.1 Geometry of the model


The geometry represents well the physical building as designed, presented in chapter 3.2.2, with just two dif-
ferences with the actual building design. Indeed, the two irregularities have been disregarded in order to present
with a report that is more representative of the general Colombian building stocks rather than of some particular
features very specific to the building itself. Vertically, it has been considered that the whole building stands
on the ground at the same level, hence ignoring the gap visible in Figure 3.4. For the in-plane regularization
of the building, the asymmetry coming from the little difference of length (0.3 m) of wall H located at Q|2 with
wall G located at G|2 has been disregarded.

52
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

5.1.2 Selected material behaviour models


The material behaviour models for steel and concrete are the same as for TW1, presented in chapter 4.2.2.2,
i.e. the Mander et al. model for concrete, and the Menegotto-Pinto model for steel rebars. The selected material
properties were the ones prescribed by the design engineer, presented in chapter 3.2.2. Namely, the (mesh)
rebars in the walls were assigned a yield resistance of fy = 490 MPa, and for the concrete, the uniaxial com-
pressive strength was set to fc’ = 35 MPa for the first floor, fc’ = 28 MPa for the second floor, and fc’ = 21 MPa
for floors 3 to 9, according to the distribution planned by the design engineer discussed in section 3.2.2. All
concrete were assigned a confinement factor of 1.15 because of the similarities of the wall reference building
with wall TW1 (e.g. transverse reinforcement ratio of ρt = 0.177% for TW1, and ρt = 0.131% for the reference
building).

5.1.3 Member modelling, floor constraints and other restraints


5.1.3.1 Walls
The walls of the reference building, as the wall TW1 presented in chapter 4, have been modelled with one
inelastic force-based frame element (infrmFB) per story, with 6 integration points (IP). The section was dis-
cretized in 300 fibres. Almost all shape of wall (Rectangular, T, L, C) could directly be modelled by elements
present in the SeismoStruct library. The only exception to that was Wall 1A4, which had a monosymmetric
double-T shape that was not available in the SeismoStruct library. The wall was hence separated in two mon-
osymmetric T-shaped walls (walls 1A/2 and 4A/2), each having a web of half the length of wall A. Those
walls were modelled with two infrmFB elements per floor, having 3 IP per elements, and were linked together
by rigid links blocking all relative movements at the end of each elements. Each longitudinal rebar in walls
was implemented separately, to the best interpretation of the structural plans. There was no difference in the
properties between the cover concrete, which had a thickness of 25 mm, and the core concrete.
5.1.3.2 Slabs
The slabs per se have not been modelled. To simulate their influence, a set of rigid links blocking translations
in directions X and Y as well as rotation around axis Z has been implemented, linking the centroid of all walls
and of the slab together in a triangulation pattern shown in Figure 5.2, hence making the walls completely
interdependent regarding horizontal movements. Figure 5.2 also indicates the wall denomination used in the
model, largely inspired by the wall numbering of the design engineer shown in Figure 3.3. The constraints
settings of the SeismoStruct model were left to their default values.

53
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Figure 5.2 – Link pattern used to simulate the slab effect

Figure 5.3 – Influence zone of each walls

54
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

5.1.4 Mass attribution


For the mass distribution, following the discussion of chapter 3.2.2, the total dead load of 6.5 kN/m2 and 25%
of the 1.8 kN/m2 (0.25·0.18 = 0.045 t/m2) live loads have been considered for the apartments floors, and 3.5
kN/m2 of dead loads and 25% of the 0.5 kN/m2 (0.25·0.5 = 0.125 t/m2) live loads have been taken into account
for the roof. In order to estimate the (lumped) mass supported by each wall, an estimation of the influence zone
of each wall has been carried out by hand. This evaluation can be seen in Figure 5.3, and the resulting mass
applied at each level of each wall, including the roof, can be seen in Table 5.1. The rotational mass provided
by the spatial repartition of punctual masses was considered sufficiently accurate, in particular since the present
study is mostly focused on the two first vibration modes, having almost no torsional component due to the
symmetry of the building.
Table 5.1 - Mass attribution for each wall modelled in SeismoStruct as a separate element
Influence Mass supported
Wall Location area Floors Roof
m2 t t
1A/2 L|1 8.54 5.94 3.10
4A/2 L|3 10.09 7.01 3.66
4B L|4 15.22 10.58 5.52
H2K S|2 6.69 4.65 2.43
J3G E|2 6.69 4.65 2.43
L C|3 1.17 0.81 0.42
M U|3 1.17 0.81 0.42
2K S|4 7.28 5.06 2.64
3J E|4 7.28 5.06 2.64
5 S|5 4.37 3.04 1.58
6 E|5 4.37 3.04 1.58
7 P|6 9.75 6.78 3.54
8 H|6 9.75 6.78 3.54
9 S|7 5.15 3.58 1.87
10 E|7 5.15 3.58 1.87
15 M|7 2.70 1.87 0.98
16 K|7 2.70 1.87 0.98
2K S|8 6.84 4.75 2.48
3J E|8 6.84 4.75 2.48
11E N|8 8.18 5.68 2.96
12F J|8 8.18 5.68 2.96
Lbis A|9 1.38 0.96 0.50
Mbis W|9 1.38 0.96 0.50
H2Kbis S|10 5.89 4.09 2.14
J3Gbis E|10 5.89 4.09 2.14
11C13 N|10 11.34 7.88 4.11
12D14 J|10 11.34 7.88 4.11
Total per story (without walls masses) 175.30 121.83 63.54

Masses per unit area (in t/m2) 0.695 0.363

55
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

5.2 Modal Analysis and Comparison with Preliminary Estimation of the


Fundamental Period
An eigenvalues analysis of the model presented in the precedent section was carried out in order to determine
the principal vibration modes of the reference building, the numerical results of the five first vibration modes
can be found in Table 5.2, and Figure 5.4 shows their deformed shapes.
Table 5.2 – Effective modal masses of the 5 first vibration modes of the reference building model
Translational Rotational
Period
Mode Direction X Direction Y Direction Z Around axis X Around axis Y Around axis Z
s ton % ton % ton % ton·m % ton·m % ton·m %
1 0.567 942 63.43% 1 0.04% 0 0.00% 14 0.02% 20186 21.53% 0 0.00%
2 0.545 1 0.04% 941 63.36% 0 0.00% 20195 30.48% 14 0.01% 65 0.12%
3 0.202 0 0.00% 0 0.01% 0 0.00% 121 0.18% 0 0.00% 40'618 73.09%
4 0.096 302 20.32% 1 0.04% 0 0.00% 21 0.03% 12253 13.07% 0 0.00%
5 0.092 1 0.04% 295 19.86% 0 0.00% 12012 18.13% 23 0.02% 13 0.02%

(a) (b) (c)

(d) (e)
Figure 5.4 – Deformed shapes of the 5 first vibration modes of the reference building model
(a) 1st mode, (b) 2nd mode, (c) 3rd mode, (d) 4th mode, (e) 5th mode, amplification: 100’000

56
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

(a) (b) (c)


Figure 5.5 – First and second vibration modes of a clamped beam in a two-dimensional space [36]
It is clearly noticeable that the two first modes are principally flexional, comparable the first vibration mode
of a cantilever beam in a two dimensional space, as illustrated in Figure 5.5 (b). While the first vibration mode,
acting in direction X, does not present any torsion due to its perfect symmetry on both sides of the Y axis, the
second vibration mode, acting in direction y has a slight torsional component (around Z) since its symmetry
in direction Y is not perfect. The third mode is clearly torsional, with no translational component at all. Finally,
the fourth and fifth modes, acting respectively in directions X and Y, are comparable to the second vibration
mode of a cantilever beam vibrating in a two-dimensional plane, as illustrated in Figure 5.5 (c).
Table 5.3 compares the two first vibration modes from the eigenvalues analysis with the estimations carried
out during the preliminary analysis with the help of simplified methods found in chapter A.4 of the NSR-10.
Logically, estimation I, based solely on the number of floor of the building gives a very gross estimation of
the period. The second estimation gives a better approximation, since it takes directly into account the total
height of the building, and most of all its coefficient depend on the structural system type (here shear walls),
and are hence fitted to correspond the construction practice. Finally, the third estimation takes into account the
actual structural system of the building through an empirical formula. However, this formula takes into account
the wall lengths in each directions, but only considers the webs of walls. It hence greatly overestimated the
fundamental period in direction X (by a factor of 1.61), since the additional stiffness provided by the many
flanges acting in that direction were not taken into account. In direction Y, the period estimation III was only
slightly underestimated
Table 5.3 – Comparison of the fundamental vibration modes estimations with the eigenvalues analysis
Direction Estimation I Estimation II Estimation III SeismoStruct
X 0.912 0.567
0.900 0.491
Y 0.451 0.545
Applying the method presented by Fajfar in his article from 1999 Capacity Spectrum Method Based on Ine-
lastic Demand Spectra [37], the modal participation factors were calculated according to eq. (5.1). The as-
sumed displacement shape corresponds to the vibration modes 1 and 2 respectively direction X and Y.
∑𝑁
𝑖=1 𝑚𝑖 Φ𝑖 1.554 for direction X (mode 1)
Γ= 𝑁 ={ (5.1)
2
∑𝑖=1 𝑚𝑖 Φ𝑖 1.553 for direction Y (mode 2)
where Γ is the modal participation factors for the fundamental modes in direction X or Y of the reference
building, N symbolizes the total number of nodes of the model, mi denotes the mass of node i, and finally Φi
stands for the displacement of node i in direction X, respectively Y, for the vibration mode 1, respectively 2,
of the reference building.
To transform the force-deformation relationship from the global reaction of the global system of the building
to the equivalent SDOF, eq. (5.2) is used, according to Fajfar [37].
𝑄 = Γ ∙ 𝑄∗ (5.2)
where Q represents the quantities in the global system (top displacement Dt, base shear V), and Q* symbolizes
the quantity in the equivalent SDOF (displacement D*, force F*).
It is to be noted that Fajfar [37] uses the definition used in eq. (5.3) to define the modal mass m*, while the
“traditional” definition of the modal mass, used in most FEM software such as SeismoStruct follow the defi-
nition shown in eq. (5.4), with the difference that the assumed displacement shape must be equal to the modal
displacement.

57
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation
𝑁
∗ 607.8 ton for direction X (mode 1)
𝑚𝐹𝑎𝑗𝑓𝑎𝑟 = ∑ 𝑚𝑖 Φ𝑖 = { (5.3)
606.7 ton for direction Y (mode 2)
𝑖=1
2 𝑁

(∑𝑁
𝑖=1 𝑚𝑖 Φ𝑖 ) 944.5 ton for direction X (mode 1)
𝑚𝐹𝐸𝑀 = = Γ ∙ ∑ 𝑚𝑖 Φ𝑖 = { (5.4)
𝑁
∑𝑖=1 𝑚𝑖 Φ𝑖2 941.8 ton for direction Y (mode 2)
𝑖=1
However, this difference in the definition of the effective modal mass does not lead to significant differences
in the final results (displacement demand), as it will be shown in the following section.

5.3 Nonlinear Static Analysis of the reference building


Nonlinear Static Analyses (NSA) are procedures that use, directly or indirectly, the response of an equivalent
Single-Degree-Of-Freedom (SDOF) system having an elastic-plastic force-displacement relationship.
In this section, the failure criterion of the model was first determined, then the capacity curve of the building
in all directions was evaluated with the help of the software SeismoStruct, to which the abovementioned failure
criterion was applied. Two different NSA procedures were then applied to assess the behaviour of the reference
building under design seismic loading: the Displacement Coefficient Method (DCM) and the Capacity Spec-
trum Method (CSM).

5.3.1 Definition of the failure criterion


Due to the similarities between wall TW1 presented in section 4.1 and the walls of the reference building
(similar thickness, single layer of reinforcement without confining reinforcement, similar transverse reinforce-
ment ratios), the experimental data coming from its test were used to define drift limits. The drift limits were
calculated for the walls of the reference building likely to lead the response of the building and its failure, i.e.
the longest walls in each direction.
For direction X, walls A, B, C, and D were considered since they were the longest in this direction. It is to be
noted that wall B is a bit shorter that the other ones (2.7 m for 3.3 m for the others), but has a free end, which
is likely to fail earlier when subjected to compression (when pushing in X positive), while all other walls are
stabilized by a flange at their extremities. When pushing in direction Y (positive and negative), walls 1, 4, 7,
8, 13 and 14 were considered. Wall 1 is significantly larger than all the other walls (5.7 m for 3.9 m for wall
4) and has no flanges at its extremities. However, the torsion due to the eccentricity of between the centre of
mass and the centre of rigidity in the Y direction result to lower inter-story drifts for wall 1 than the other
walls, making it necessary to check which wall reaches its limit first. All walls used to determine the drift
limits have been filled in with grey in Figure 3.3.
As it can be observed in Figure 4.14, wall TW1 failed when pushed in direction positive, i.e. when its free end
was compressed. The failure occurred after two complete cycles of drifts of 16.3 mm displacement, followed
by half a cycle of displacement of 22.2 mm when pushing towards the flange, i.e. the negative direction. The
inter-story drift limits corresponding to the state of Near Collapse (NC, as defined by the EN 1998-3 [34]) of
TW1 are hence supposed in this paper as ΔNC = 0.73% for the positive direction, and ΔNC = 1.00% for the
negative direction. Those limits correspond to 16.3 mm of top displacement for the positive direction, and
22.2 mm of top displacement in the negative direction.
The Near Collapse chord rotation of each wall (TW1, A, B, C, D, 1, 4, 7, 8, 13 and 14) were calculated
according to eq. (5.5), corresponding to eq. A.1 of the appendix A of the EN 1998-3, as found in the 2013
corrigendum [35]. The results can be found in Table 5.4 and Table 5.5.
0.225 𝑓
1 max(0.01; 𝜔′ ) ′ 𝐿𝑠 0.35 𝑘𝑐 𝜌𝑡 𝑓𝑦𝑡′
𝜃𝑁𝐶 = 0.016 ∙ 0.3𝜈 [ 𝑓] (min (9; )) 25 𝑐 1.25100𝜌𝑑 (5.5)
𝛾𝑒𝑙 max(0.01; 𝜔) 𝑐 ℎ𝑤
where θNC stands for the chord rotation corresponding to the Near Collapse (NC) limit state; γel is a factor equal
to 1.5 for primary and 1.0 for secondary seismic elements, ν represents the axial load ratio, ω and ω’ are
respectively the mechanical reinforcement ratio of the tension and compression longitudinal reinforcement, Ls
denotes the shear span, calculated through eq. (5.6) which was adapted from eq. (3.35) of Priestley et al. 2007
[25], lw symbolizes the wall length, kc is the confinement effectiveness factor (here taken as 1.15), ρt depicts
the transversal reinforcement ratio, fyt is the stirrup rebar yield strength (here taken as the transverse rebar yield
58
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

strength), fc’ stands for the concrete cylinder strength, and ρd denotes the reinforcement ratio of diagonal rein-
forcement (here equal to 0).
∑𝑁
𝑖=1 𝑚𝑖 Φ𝑖 ℎ𝑖 15.88 m for direction X (mode 1)
𝐿𝑠 = ={ (5.6)
∑𝑁 𝑚
𝑖=1 𝑖 𝑖Φ 15.89 m for direction Y (mode 2)
where N is the number of control nodes, mi symbolizes the nodal mass of node i, Φi denotes the assumed
displacement shape, equal to the fundamental vibration modes, and hi stands for the altitude from the base of
node i.
The failure criterion chosen for the response of the reference building to the design earthquakes corresponds
to the Significant Damage (SD) limit state found in the EN 1998-3, adapted with the experimental data from
the testing of TW1. According to article A.3.2.3 of the EN 1998-3, the chord rotation corresponding to the
Significant Damage limit state is equal to θSD = ¾θNC. The failure inter-story drifts for the first floor were
calculated according to eqs. (5.7) to (5.9).
𝜃𝑁𝐶,𝑖
Δ𝑁𝐶,𝑖 = Δ𝑁𝐶,𝑇𝑊1 ∙ (5.7)
𝜃𝑁𝐶,𝑇𝑊1
δ𝑁𝐶,𝑖 = Δ𝑁𝐶,𝑖 ∙ ℎ𝑝 (5.8)

δ𝑆𝐷,𝑖 = 3⁄4 δ𝑁𝐶,𝑖 Δ𝑆𝐷,𝑖 = 3⁄4 Δ𝑁𝐶,𝑖 𝜃𝑆𝐷,𝑖 = 3⁄4 𝜃𝑁𝐶,𝑖 (5.9)

where i stands for the studied wall (1, A, B, C/D), θNC is the chord rotation at the Near Collapse limit state of
the wall, ΔNC and ΔSD denote the inter-story drift corresponding respectively to the Near Collapse (SD) and
Significant Damage (SD) limit states, hp symbolizes the inter-story height, and δNC and δSD are the relative
displacement for the first story corresponding respectively to the Near Collapse (SD) and Significant Damage
(SD) limit states.
The fitting with the results of the experimental data coming from wall TW1 was done using the drift limit for
TW1 in the positive direction for wall B when pushing in direction X positive, and for wall 1, 4, 7, 8, 13 and
14 since they did not present a flange at their compressed extremities. For walls A, B when pushing in the X
negative direction, C, and D, the drift limit for TW1 in the negative direction was used. The results can be
found in Table 5.4 and Table 5.5. The axial load ratios for the walls of the reference building were calculated
with the outputs of the SeismoStruct analysis. The mechanical reinforcement ratios of TW1 were taken as the
average mechanical reinforcement ratios of the boundary elements of the walls, excluding the flange as the
empirical formula was most likely fitted with experimental results on symmetric walls.

59
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Table 5.4 – Calculation of the drift limit for the walls of the reference building (direction X)
Parameter TW1 (direction +) TW1 (direction -) B (direction X+) B (direction X-) A (direction X±) C/D (direction X±)
γ el - 1.5 1.5 1.5 1.5 1.5 1.5
ν - 5.0% 5.0% 5.3% 5.3% 3.9% 3.8%
ω - 0.481 0.519 0.018 0.018 0.018 0.018
ω’ - 0.519 0.481 0.018 0.018 0.018 0.018
Ls m 10.0 10.0 15.9 15.9 15.9 15.9
lw m 2.7 2.7 2.7 2.7 3.3 3.3
kc - 1.15 1.15 1.15 1.15 1.15 1.15
ρt - 0.177% 0.177% 0.131% 0.131% 0.131% 0.131%
f yt MPa 460 460 490 490 490 490
f c ’ MPa 28.4 28.4 35 35 35 35
ρd - 0% 0% 0% 0% 0% 0%
hp m 2.21 2.21 2.40 2.40 2.40 2.40
θ NC rad 0.0382 0.0369 0.0443 0.0443 0.0420 0.0421
ΔNC - 0.74% 1.00% 0.86% 1.21% 1.15% 1.15%
δ NC mm 16.3 22.2 20.6 29.0 27.5 27.5
θ SD rad 0.0286 0.0276 0.0332 0.0332 0.0315 0.0316
ΔSD - 0.55% 0.75% 0.64% 0.91% 0.86% 0.86%
δ SD mm 12.2 16.7 15.4 21.7 20.6 20.6
Table 5.5 – Calculation of the drift limit for the walls of the reference building (direction Y)
Parameter TW1 (direction +) TW1 (direction -) 1 (direction Y±) 13/14 (direction Y±) 4 (direction Y±) 7/8(direction Y±)
γ el - 1.5 1.5 1.5 1.5 1.5 1.5
ν - 5.0% 5.0% 3.9% 3.8% 5.3% 5.8%
ω - 0.481 0.519 0.018 0.018 0.018 0.018
ω’ - 0.519 0.481 0.018 0.018 0.018 0.018
Ls m 10.0 10.0 15.9 15.9 15.9 15.9
lw m 2.7 2.7 5.7 2.9 3.9 3.7
kc - 1.15 1.15 1.15 1.15 1.15 1.15
ρt - 0.177% 0.177% 0.131% 0.131% 0.131% 0.131%
f yt MPa 460 460 490 490 490 490
f c ’ MPa 28.4 28.4 35 35 35 35
ρd - 0% 0% 0% 0% 0% 0%
hp m 2.21 2.21 2.40 2.40 2.40 2.40
θ NC rad 0.0382 0.0369 0.0347 0.0440 0.0390 0.0395
ΔNC - 0.74% 1.00% 0.67% 0.85% 0.75% 0.76%
δ NC mm 16.3 22.2 16.1 20.4 18.1 18.3
θ SD rad 0.0286 0.0276 0.0260 0.0330 0.0292 0.0296
ΔSD - 0.55% 0.75% 0.50% 0.64% 0.57% 0.57%
δ SD mm 12.2 16.7 12.1 15.3 13.6 13.7
It is noticeable that the most restrictive wall in direction X positive is wall B, reaching the Significant Damage
(SD) limit state at an inter-story drift of 0.64% (15.4 mm of relative lateral displacement), and the Near Col-
lapse (NC) limit state at an inter-story drift of 0.86% (20.6 mm of relative lateral displacement). In direction
X negative, the critical wall is wall A, reaching the SD limit state at 0.86% inter-story drift (20.6 mm of relative
lateral displacement) and the NC limit state at 1.15% inter-story drift (27.5 mm of relative lateral displace-
ment). In direction Y positive and negative, wall 1 reaches the SD limit state at 0.50% inter-story drift
(12.1 mm of relative lateral displacement), and the NC limit state at 0.67% inter-story drift (16.1 mm of relative
lateral displacement). A verification of the evolution of the inter-story drifts for each wall through the pushover
analysis presented in section 5.3.2 showed that wall 1 was the first to reach both limit states, despite the tor-
sional effects leading to lower inter-story drifts for wall 1 than other walls.
It should be noted that the Significant Damage limit state drift limits calculated here are lower than the drift
limits set by the NSR-10 for buildings with reinforced concrete walls presented in section 2.1.1.10. Those drift
limits are however meant to ensure a good behaviour at the serviceability limit state of the building and are

60
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

calculated via an elastic linear analysis, while the limits calculated here are applicable for a highly nonlinear
analysis, and are meant to ensure the structural safety of the building. They are hence not directly comparable.

5.3.2 Pushover analysis and application of the failure criterion


During the pushover analyses, the reference building response has been studied when pushed in each orthog-
onal direction separately (X+, X-, Y+, and Y-). In every analysis, the node at the centre of mass of the top slab
was gradually loaded by an incremental lateral displacement of 1 mm for each load step. The analyses were
response controlled and were run until the model could not converge towards a solution for the latest iteration.
The convergence criteria as well as the iterative strategy settings were left to their default values. The axial
loads were automatically calculated by SeismoStruct from the implemented masses and self-weight of the
walls, without need to explicitly implement them in the model.
The influence of the accidental torsion as proposed by section 3.2.2.2 of the FEMA-356 [22] (eccentricity of
the displacement application point of 5% of the slab dimension) was studied. It was verified that the accidental
torsion influenced the lateral displacement by less than 10% of the total lateral displacement, and was hence
not considered in the present analyses, in accordance with article 3.2.2.2.2.3 of the FEMA-356 [22].
The top displacement at which the drift limits determined in section 5.3.1 were reached in each direction can
be observed in Table 5.6. Figure 5.6 shows the global response of the reference building in each direction. The
dotted lines indicates the response after the Significant Damage limit state is reached, and the interrupted line
is used to plot the response after the Near Collapse limit state is exceeded, until the model could not reach
convergence anymore.
Table 5.6 – Top lateral displacement at SD and NC limit states of the reference building for each direction
Top lateral displacement
Limit state Direction X+ Direction X- Direction Y+ Direction Y-
mm mm mm mm
Significant Damage 414 512 409 408
Near Collapse 464 583 447 445
It should be noted that the software SeismoStruct does not consider the shear resistance of walls. However,
some quick conservative verifications of the wall shear capacity revealed that the shear resistance of the main
walls of the reference was sufficient to sustain the shear force associated with the development of the moment
resistance of said walls.
Due to the symmetry of the building, the response in direction Y positive and negative are logically almost
perfectly equal. In direction X, while most of the response in negative and positive directions is very similar,
the Significant Damage and Near Collapse limit states are reached in the positive direction significantly before
the negative direction. For these reasons, as well as for the sake of clarity and readability, only the response in
positive directions (X and Y) will be further discussed in the present report.

61
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Global response of the reference building


1200
NC: 447;1149
1100
SD: 408;1170 NC: 445;1146
1000
900 SD: 409; 1169

800 NC: 464;561


Base shear [kN]

700 SD: 414;614


SD: 512;573 X+
600
X-
500
Y+
NC: 583;550
400 Y-
300
200
100
0
0 50 100 150 200 250 300 350 400 450 500 550 600 650
Lateral displacement of the top slab centroid [mm]

Figure 5.6 – Global response of the reference building

5.3.3 Behaviour of the reference building under design seismic loading


In this section, the behaviour of the reference building when subjected to the design earthquake is assessed.
This was carried out with the Displacement Coefficient Method (DCM), as implemented in the appendix A-3
of the NSR-10 [2] (herein explained in section 2.3), and the Capacity Spectrum Method (CSM) as proposed
by Fajfar [37]. The DCM includes directly a response idealization procedure considering post-yield strain
hardening. The CSM however demands for the response idealization to be elastic-perfectly plastic, and does
not directly impose an idealization procedure, but rather gives some recommendations while leaving the choice
to the engineer. Because of the difference in the required post-yield idealized response for both methods, dif-
ferent idealizations were used to assess the displacement demand with the DCM and the CSM.
5.3.3.1 Displacement Coefficient Method
For the displacement coefficient method, two elastic-plastic bilinear idealizations procedures (with post-yield
strain-hardening) were considered: EPSH1 and EPSH2. The first idealization (EPSH1) follows the procedure
proposed by appendix A-3 of the NSR-10 [2], herein presented in section 2.3. The second (EPSH2) varies
slightly. Both procedures are detailed below.
Three principles were applied to define the EPSH1 idealization:
 The first segment of the idealized curve crosses the original capacity curve at a shear level of 0.6Vy,
where Vy is the idealized yield base shear.
 The second segment of the idealized curve crosses the original capacity curve at displacement level
δT, where δT is the target displacement calculated through eq. (2.38) repeated here in eq. (5.10).
𝑇𝑒 2 (5.10)
𝛿𝑇 = 𝐶𝑜 𝐶1 𝑆𝑎𝑒 ( ) 𝑔
2𝜋
Where Co and C1 are factors calculated according to eqs. (2.39) and (2.40), Sae is the elastic demand of the
equivalent SDOF, Te is the effective period of the building and g is the earth standard acceleration
 The areas below the original capacity curve and the idealized one are equivalent between a displace-
ment of zero and δT. (Equal energy principle)
For idealization EPSH2, the principles were slightly modified:
 The first segment of the idealized curve crosses the original capacity curve at a shear level of 0.75Vy.
62
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

 The second segment of the idealized curve crosses the original capacity curve at displacement level
2δT, where δT is the target displacement calculated through eq. (5.10).
 The equal energy principle was not respected, and the idealized yield base shear was left to the best
judgement of the author.
Some remarks should be made regarding the calculation of the effective period Te, target displacement δT, and
other parameters:
 The effective period Te of the building was not calculated with eq. (2.37), but rather with eq. (5.11)
below.

𝑚𝐹𝐸𝑀
𝑇𝑒 = 2π√ (5.11)
𝐾𝑒
𝑉𝑦
𝐾𝑒 = (5.12)
𝛿𝑦
where m*FEM is the effective modal mass calculated through eq. (5.4), Ke symbolizes the effective
stiffness of the building, and δy stands for the effective yield displacement of the control point.
 Factor C1 in eq. (5.10) for the calculation of the target displacement is in fact the modal participation
factor Γ calculated in eq. (5.1).
 The acceleration spectrum used in eq. (5.10) is the spectrum coming from the microzoning campaign
presented in section 3.2.2.
 The ductility factor Rd was calculated with eq. (2.41) proposed by the NSR-10 [2] using the global
building mass M, and not with the definition given by FEMA-356 [22] (shown in eq. (5.13)) that
considers the effective modal mass m*FEM. The total building mass M considered in the analysis is the
“activable” mass; the sum of all nodal masses for nodes whose movements were not restrained (i.e. all
nodes but the base ones).
𝑆𝑎𝑒 ∗
𝑅𝑑 = 𝑚 (5.13)
𝑉𝑦 𝐹𝐸𝑀
where Rd is the ductility factor calculated according to the FEMA-356 [22], Sae is the elastic accelera-
tion demand for a SDOF of period Te, Vy is the idealized yield base shear, and m*FEM is the effective
modal mass calculated through eq. (5.3).
This definition was chosen because the definition given by the FEMA-356 mixes the quantities of the
SDOF (Sae, m*FEM) and the quantities of the MDOF (Vy). This seems to lead to an inconsistency
 The ductility demand was directly calculated through eq. (5.14)
𝛿𝑇
𝜇= (5.14)
𝛿𝑦
where δT is the target displacement calculated through eq. (5.10) and δy stands for the effective yield
displacement of the control point.
 The safety factor γSF was calculated with eq. (5.15).
𝐷𝑆𝐷
𝛾𝑆𝐹 = (5.15)
𝛿𝑇
Where γSF is the safety factor, and DSD is the top-story drift at which the Significant Damage limit state is
reached, shown in Table 5.6.
The idealizations shown in Figure 5.7 for direction X+, and Figure 5.8 for direction Y+ are the response of the
global MDOF. Points YP1 and YP2 represent the yielding points of respectively idealization EPSH1 and
EPSH2, TD1 and TD2 are the points of the original capacity curve at which the target displacement is reached,
points SD1 and SD2 correspond to the Significant Damage limit state on the idealized response EPSH1 and
EPSH2, while NC1 and NC2 relate to the Near Collapse limit state. And finally 2·TD2 is the point at which
the post-yield segment of EPSH2 crosses the original capacity curve, at a displacement of 2δT. The response

63
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

of the equivalent SDOF can be found by dividing displacements and forces by the corresponding modal par-
ticipation factor Γ. The detailed results of the analysis can be found in Table 5.7. Both idealizations in both
directions lead to displacement demands significantly smaller than the failure criterion considered, the SD
limit state, with a safety factor γSF varying from 2.5 and 2.6 for EPSH2 to 3.1 for EPSH1.

Figure 5.7 – EPSH1 and EPSH2 idealizations of the building response in direction X+

Figure 5.8 – EPSH1 and EPSH2 idealizations of the building response in direction Y+

64
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Table 5.7 – Detailed results of the Displacement Coefficient Method analysis


Direction Direction X+ Direction Y+
Idealization EPSH1 EPSH2 EPSH1 EPSH2
Vy kN 399 500 680 900
δy mm 9.80 25.15 15.70 37.97
*
m ton 944.5 944.5 941.8 941.8
M ton 1484.7 1484.7 1484.7 1484.7
Ke kN/mm 40.70 19.88 43.30 23.70
Te s 0.96 1.37 0.93 1.25
S ae m/s2 3.68 2.28 3.85 2.57
Sd mm 85.5 108.3 83.7 102.1
δT mm 132.8 168.3 129.9 158.5
D SD mm 414 414 409 409
Rd - 13.7 6.8 8.4 4.2
μ - 13.5 6.7 8.3 4.2
γ SF - 3.1 2.5 3.1 2.6
Regarding the idealization proposed by the NSR-10 (EPSH1), it appears that the procedure leads, for the ref-
erence building, to important estimations of the effective elastic stiffness Ke and post-yield residual stiffness
(strain-hardening). While the important evaluation of the effective elastic stiffness Ke induces a high elastic
acceleration demand (and hence high elastic base shear demand), the important post-yield strain-hardening
leads to a significantly low estimation of the effective yield base shear Vy (and displacement δy). Those two
effects combined induced very high evaluations of the ductility factor and ductility demand (Rd = 13.7 and
μ = 13.5), while the displacement demand is actually not very important.
It is notable that the important post-yield strain-hardening of the EPSH1 idealization leads to very important
overestimations of the residual strength at high level of displacement, after the target displacement (TD1) is
exceeded. In itself, this is not a problem since the analysis does not take this part of the idealized response in
the calculations. However, this reveals a problem linked to the DCM as implemented by the NSR-10 (and the
FEMA-356): The evaluation of the building requires each time a new idealization of the response for different
solicitations (e.g. damage control, ULS, etc.). While not intrinsically problematic, this could easily lead to
confusion in the interpretation of results and mislead conclusions, not to mention the risk of oversight by the
engineer could lead to errors in the determination of the displacement demand.
Idealization EPSH2 logically resulted in slightly larger displacement demands due to its lower initial
stiffness Ke. Paradoxically, its lead to lower (and slightly more realistic) ductility demand and ductility factors
due to its higher evaluation of the yield displacement δy and base shear Vy. Moreover, the evaluations of the
residual strength one the target displacement (TD2) is exceeded correspond better to the original capacity curve
calculated with SeismoStruct.
Finally, there is a small difference between the evaluations of the ductility factor Rd and the ductility demand
μ, while theoretically those two parameters should be equal for buildings having periods greater than TC. The
difference is of 1.14% in direction X+ and 1.55% in direction Y+ (For both idealization procedures). The
author supposes that this difference comes from a problem of normalization of the assumed shape building,
leading to inconsistencies between the modal participation factor, the effective mass and the total building
mass. Such little difference have however no significant importance for engineering applications and has been
disregarded in the rest of this report.
5.3.3.2 Capacity Spectrum Method
The capacity spectrum method, as proposed by Fajfar [37], was also used to calculate the displacement demand
of the reference building. This method requires for the bilinear idealization to be elastic-perfectly plastic be-
cause of the definition of the reduction factor due to ductility Rd, but no other explicit requirements are stated
in the paper. It is left to the engineer to judge what the proper idealization is, and only a suggestion is made to
cross the original capacity curve at 0.6Vy. Two elastic-perfectly plastic idealizations of the capacity curve were
considered: EPP1 and EPP2. Both procedures are detailed below.
Three principles were applied to define the EPP1 idealization:
65
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

 The first segment of the idealized curve crosses the original capacity curve at a shear level of 0.6Vy,
where Vy is the idealized yield base shear.
 The second segment of the idealized curve has a stiffness equal to zero.
 The areas below the original capacity curve and the idealized one are equivalent between a displace-
ment of zero and δT, where δT is the target displacement calculated through eq. (5.10).
For idealization EPP2, the first principles was slightly modified:
 The first segment of the idealized curve crosses the original capacity curve at a shear level of 0.75Vy,
where Vy is the idealized yield base shear.
 The second segment of the idealized curve has a stiffness equal to zero.
 The areas below the original capacity curve and the idealized one are equivalent between a displace-
ment of zero and δT.
It is important to note that, in a spirit of consistency, the definition of the effective modal mass used in the
effective period calculation was not m*FEM as in eq. (5.4), but rather the definition given by Fajfar [37], illus-
trated in eq. (5.3). The effective period Te is hence not calculated with eq. (5.11) as for the DCM, but with the
slightly modified eq. (5.16).

𝑚𝐹𝑎𝑗𝑓𝑎𝑟
𝑇𝑒 = 2π√ (5.16)
𝐾𝑒

where m*Fajfar is the effective modal mass calculated through eq. (5.3), and Ke symbolizes the effective stiffness
of the building calculated through (5.12).
Both idealizations are shown in Figure 5.9 for direction X+, and Figure 5.10 for direction Y+. Note that the
response presented here is the answer of the global MDOF system, and the response of the equivalent SDOF
system can be found by dividing displacements and forces by the corresponding modal participation factor Γ,
as shown in eq. (5.17). Points YP1 and YP2 represent the yielding points of respectively idealization EPP1
and EPP2, TD1 and TD2 are the points of the original capacity curve at which the target displacement is
reached, points SD1 and SD2 correspond to the Significant Damage limit state on the idealized response EPP1
and EPP2, while NC1 and NC2 relate to the Near Collapse limit state.
In order to use the CSM, the response of the equivalent SDOF was calculated through eq. (5.17).
𝑉 𝐷𝑡
𝐹∗ = and 𝐷∗ =
Γ Γ
and especially (5.17)
𝑉𝑦 𝛿𝑦
𝐹𝑦∗ = and 𝐷𝑦∗ =
Γ Γ
where V is the base shear in the global MDOF, Dt stands for the displacement of the top control point in the
global MDOF, Γ symbolizes the modal participation factor, F* represents the force applied on the equivalent
SDOF, D* denotes the displacement of the equivalent SDOF, Vy and δy depict respectively the idealized base
shear and top displacement of the global MDOF at the yielding point, and F*y and D*y are the yield force and
displacement of the equivalent SDOF.

66
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Figure 5.9 – EPP1 and EPP2 idealizations of the building response in direction X+

Figure 5.10 – EPP1 and EPP2 idealizations of the building response in direction Y+
The idealization EPP1, following the suggestion of Fajfar [37] to cross the original capacity curve at 0.6Vy
combined with the equal energy principle, lead to a very high evaluation of the effective stiffness Ke and low
estimation of the effective yield base shear Vy. This will be further discussed below.
Since the graphical procedure of the CSM uses the inelastic Acceleration-Displacement Response Spectrum
(ADRS), the ADRS for each direction and idealization has been calculated through eqs. (5.18) to (5.23). In
order to simplify the analysis, the conservative assumption of eq. (6a) from Fajfar [37], presented here in eq.
(5.20) was made.
𝐹𝑦∗
𝑆𝑎𝑦 = ∗ ≤ 𝑆𝑎𝑒 (5.18)
𝑚𝐹𝑎𝑗𝑓𝑎𝑟 ∙𝑔

67
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

𝑆𝑎𝑒
𝑅𝑑 = (5.19)
𝑆𝑎𝑦
𝑇𝑜′ = 𝑇𝐶 (5.20)
𝑇𝑜′
(𝑅 − 1) + 1 if 𝑇𝑒 < 𝑇𝑜′
𝜇Δ = { 𝑑 𝑇𝑒 (5.21)
𝑅𝑑 if 𝑇𝑒 ≥ 𝑇𝑜′
𝑆𝑎𝑒
𝑆𝑎 = (5.22)
𝑅𝑑
𝜇 𝜇 𝑇2 𝑇2
𝑆𝑑 = 𝑆𝑑𝑒 = 𝑆 = 𝜇 𝑆 (5.23)
𝑅𝑑 𝑅𝑑 4𝜋 2 𝑎𝑒 4𝜋 2 𝑎
where Say is the spectral acceleration at the yielding point, Sae stands for the elastic spectral acceleration for the
idealized SDOF of period Te, g symbolizes the earth standard acceleration. Rd denotes the reduction factor due
to ductility (ductility factor), TC is the period corresponding to the end of the acceleration spectrum plateau
and To’ depicts the period corresponding to the beginning of the constant ductility factor (Rd), μΔ stands for the
displacement ductility, Sa represents the inelastic acceleration spectrum, Sd is the inelastic displacement spec-
trum, Sde denotes the elastic displacement spectrum, and T stands for the period.
The graphical Acceleration-Displacement Response Spectrum (ADRS) can be observed in Figure 5.11 for
direction X+ and Figure 5.12 for direction Y+. Every radial line passing through the origin corresponds to the
elastic response of a SDOF system of a definite period, in our case, the SDOF of effective period Te corre-
sponding to the evaluation from idealizations EPP1 and EPP2. Points YP1 and YP2 represent the spectral
demand at yielding point of respectively idealization EPP1 and EPP2. ED1 and ED2 show their elastic spectral
demand, and PP1 and PP2 are the performance point (i.e. the inelastic demand) of the reference building,
according to idealizations EPP1 and EPP2. Points SD and NC show the Significant Damage and Near Collapse
limit states of the building, translated in the equivalent SODF system. The detailed results of the CSM analysis
can be found in table Table 5.8. The safety factor γSF has been calculated through eq. (5.15) as for the DCM,
since the target displacement δT calculated through eq. (5.10) corresponds exactly to the translation to the
MDOF of the displacement of the performance point.

Figure 5.11 – Acceleration-Displacement Response Spectrum of the equivalent SDOF in direction X+

68
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

Figure 5.12 – Acceleration-Displacement Response Spectrum of the equivalent SDOF in direction Y+


Table 5.8 – Detailed results of the Capacity Spectrum Method analysis
Direction Direction X+ Direction Y+
Idealization EPP1 EPP2 EPP1 EPP2
Vy kN 493 557 832 945
δy mm 15.06 37.05 22.09 45.02
*
m ton 607.8 607.8 606.7 606.7
M ton 1484.7 1484.7 1484.7 1484.7
Ke kN/mm 32.74 15.04 37.67 20.99
Te s 0.86 1.26 0.80 1.07
S ae m/s2 4.28 2.54 4.70 3.18
Sd mm 79.4 102.6 75.7 91.9
δT mm 123.4 159.5 117.6 142.6
D SD mm 414 414 409 409
Rd - 8.2 4.3 5.3 3.2
μ - 8.2 4.3 5.3 3.2
γ SF - 3.4 2.6 3.5 2.9
As for the DCM, both idealizations in both directions lead to displacement demands significantly smaller than
the failure criterion considered, the SD limit state, with a safety factor γSF varying from 2.6 and 2.9 for EPP2
to 3.4 and 3.5 for EPP2. The high evaluation of the effective stiffness Ke and low estimation of the yield base
shear Vy by the idealization procedure EPP1 lead to important values for the ductility factor Rd and ductility
demand μ.
Due to its lower effective stiffness Ke evaluation, idealization EPP2 lead to a larger effective period Te estima-
tion and larger displacement demand δT. However, the evaluation of the ductility factor Rd and ductility demand
μ with idealization EPP2 were lower than for EPP1 due to the higher yield base shear Vy and displacement δy.
Those two (Rd and μ) were more in the range of expected values when observing the position of the target
displacement δT on the capacity curve in Figure 5.9 and Figure 5.10.
Finally, the inconsistency between Rd and μ observed in the application of the DCM in section 5.3.3.1 is not
present in the CSM since all quantities used where in the same system (the equivalent SDOF).

69
Colombian RC wall buildings assessment –Numerical analysis: Model Presentation and Results Interpretation

5.4 Comparison of the of the Nonlinear Static Analyses, Conclusions, and


Comments on the Validity of the Preliminary Analysis
The Nonlinear Static Analyses carried out with both methods all determined that the capacity of the reference
building is sufficient to withstand design earthquakes with an acceptable margin. The safety factors determined
varied from 2.6 to 3.5 depending on the analysis, the idealization of the building response, and the loading
direction considered, however most of the variation came from the selected idealization procedure.
It comes out of the analyses that both methods (DCM and CSM) lead to relatively similar displacement demand
and safety factors. The idealization chosen also seems not to change drastically the displacement demand, and
the failure criterion was by far not reached with any of the considered idealization and method.
However, the estimation of the standard engineering quantities (ductility factor Rd and ductility demand μ)
varied greatly, from 3.2 to as much as 13.7. This is concerning since those quantities are usually at the base of
the procedures promoted by construction codes (NSR-10 [2], ASCE 7 [3], EN 1998-1 [20], SIA 261 [6]), and
the blind use of a calculated value for those parameters could lead to severely mislead conclusions. The ideal-
ization procedure proposed by the NSR-10 and the FEMA-356 [22] in particular lead to extremely high eval-
uations of the ductility factor Rd (13.7 for direction X+ and 8.4 for direction Y+). This raises questions regard-
ing the validity of this method, imported from the U.S. standards, for the evaluation of constructions designed
according to typical Colombian practices.
For the analysed building, the high amount of large walls, having different lengths, lead to a very gradual
transition from the elastic to the plastic regime. It would appear that the slow evolution of the capacity curve
combined with a relatively high initial stiffness leading to low displacement demand does not correspond to
the typical situations the CSM method as implemented in the NSR-10 was originally fitted to.
Because no exact idealization was prescribed for the CSM method, it has been easier to adapt the chosen
idealization (EPP2) to the conditions of the reference building for this method in order to obtain more repre-
sentative results for the standard engineering parameters Rd and μ. While convenient for the engineer, this is
not necessarily adequate for construction codes, supposed to standardize the evaluation and design procedures
in order to ensure the good behaviour of the building stock.
In the preliminary analysis, it was determined that the ductility demand would be extremely low for the refer-
ence building, with a maximum of 1.29 with the least favourable assumptions. Moreover, when using the
preliminary analysis procedure and applying the response spectrum coming from the microzoning campaign,
the displacement demand was estimated to be lower than the actual yield displacement. This is inconsistent
with the results of the Nonlinear Static Analysis which lead to the ductility demands of 4.3 for direction X+
and 3.2 for direction Y+ in the most favourable evaluations. This difference has two origins, if not more: (i)
the yield displacements estimated in the preliminary analysis were significantly lower than the ones considered
in the Nonlinear Static Analysis, and (ii) the periods considered in the preliminary analysis were estimations
of the elastic period and hence lead to lower displacement demands than the effective period considered in the
Nonlinear Static Analysis.
In the preliminary analysis, the yield lateral displacements of the equivalent SDOF were estimated considering
the yield curvatures proposed by Prietley et al. [25]. Using the more refined estimation of the SDOF system
effective height and transferring the results to the global MDOF, they result in lateral displacement of 145 mm
for direction X and 96 mm for direction Y. These displacements are significantly larger than the yield dis-
placement considered in the Nonlinear Static Analyses (9.8 to 37. 0 mm for direction X+ and 15.7 to 45.0 mm
for direction Y+). Moreover, when placed on the capacity curve presented in Figure 5.6, it is directly noticeable
that significant inelastic deformations occur before those displacement levels. The preliminary estimations of
yield displacement hence seem not to be adapted for the reference building.

70
Colombian RC wall buildings assessment – Summary and Conclusion

6 SUMMARY OF THE THESIS AND CONCLUSIONS

In the first part of the present report, an evaluation of the Colombian design and assessment code (NSR-10)
was carried out, focusing on the requirements related to the seismic behaviour of buildings composed of rein-
forced concrete walls. The Colombian code is largely inspired by the U.S. standards, and very few adaptations
have been made to the analysis procedures and requirements. This can be a source of potential problems,
especially since construction practices in Colombia are very different to those in the U.S. The main problematic
issues identified in the Colombian code are repeated below.
The first deficiency noted in the NSR-10 is the lack of minimum thickness applicable for all reinforced con-
crete walls, regardless of the method used to design them. Combined with the absence of maximum axial load
ratio and the common use of a single layer of reinforcement sanctioned by the NSR-10, it can lead to potentially
important out-of-plane instabilities that severely reduce severely the required displacement capacity to sustain
major seismic events.
Secondly, even for structures belonging to the highest energy dissipation category, it is not specifically re-
quired for the shear resistance of walls to exceed the shear load corresponding to the development of the
flexural resistance. This can lead to fragile ruptures and is hence of special concern, especially since engineers
occasionally use beam-elements models to assess the seismic behaviour of buildings, models that typically do
not consider the shear strength of elements.
Thirdly, the formulation of the Colombian code for both the design and evaluation of structures is largely based
on the use of the equivalent lateral load method, i.e. a linear elastic analysis, and in some cases on a dynamic
linear analysis. The code also allows engineers to use nonlinear methods that give a significantly better appre-
ciation of the building capacity. However, those are mostly not directly included in the NSR-10 and are simply
mentioned as external reference. This does not efficiently promote the use of such advanced methods, espe-
cially given that those reference are all in English and the knowledge of the English language varies greatly
within the Colombian population.
The last comment on the Colombian code is related to the proposed procedure for the evaluation of existing
structures. It is based on two global indexes calculated through a linear elastic force-based analysis; the over-
stress index comparing the internal loads of all elements with their strength, and the flexibility index comparing
the inter-story drifts with flat-rate limits. This linear elastic force-based method does not take the favourable
plastic redistributions into account and does not assess the actual displacement capacity of the building. More-
over, the safety factors for the calculation of elements strengths in design do not seem to be considered and
only two global and highly subjective safety factors for the whole building are used. This does not guarantee
a sufficient safety level of the structure, as fragile failure modes are not necessarily prevented.
In the second part of the study, a simplified analysis procedure was applied to a set of 6 buildings in order to
select the reference building analysed with a numerical analysis. The results of the detailed analysis mentioned
below were then used to assess the validity of the preliminary analysis. It appeared that the simplified formulas
proposed in the Colombian code to assess the fundamental period of a building are suitable to assess the linear
elastic period of a building with rectangular walls. However, they do not consider the additional stiffness
provided by flanges and is hence not adapted to assess the period of buildings with important flanged walls.
Moreover, the formulas only give an appreciation of the linear elastic period of the building, and do not ade-
quately represent the effective period after the onset of inelastic deformations. This leads to underestimated
displacement demands. The preliminary analysis also resulted in important yield displacements evaluations.
The detailed analysis of the global building showed that inelastic deformation occurred long before those levels
of displacement were reached. Those combined effects lead to severe underestimations of the ductility demand.
The comparison of numerical and experimental results of the monotonic lateral loading of a single wall was
then used to validate the selected modelling options. It resulted that SeismoStruct, the software used to model
the reference building, provided a suitable evaluation of the response at the global level. However, the strain
predictions were not satisfactory to define a reliable failure criterion. It was therefore decided to evaluate drift
limits for the walls of the reference building in order to define a failure criterion. Those drift limits were
adapted from the experimental data with the help of an empirical formula coming from the European standards.
The calculated inter-story drift limits were significantly lower than the flat rate considered by the Colombian
71
Colombian RC wall buildings assessment – Summary and Conclusion

code (0.64% for direction X and 0.50% for direction Y, while the NSR-10 proposes a limit of 1%). However,
the calculated drift limit correspond the Ultimate Limit State (ULS) and the limits set in the NSR-10 correspond
to the Serviceability Limit State (SLS), which makes them not directly comparable.
In the final part of the study, a numerical model of the complete reference building was used to proceed to a
modal analysis, and a pushover analysis. The behaviour of the reference building under design seismic loading
was then assessed from the result of the pushover analysis with two nonlinear static analysis procedures; the
Displacement Coefficient Method (DCM) and the Capacity Spectrum Method (CSM).
It came out of these analyses that the selected reference building presented sufficient displacement capacity to
sustain the design earthquake. Although reassuring, this conclusion cannot obviously be generalized to the
Colombian building stock as further studies on other buildings would be required. It should also be noted that
the potential effects of wall out-of-plane instability and biaxial loading on the displacement capacity were not
directly considered in this study. Only the out-of-plane instabilities were indirectly taken into account since
the wall test on which the failure criterion was scaled failed after the onset of noticeable out-of-plane move-
ments. In order to assess the capacity of the reference building, a more thorough analysis would be needed.
Such analysis would typically include instabilities evaluation and a nonlinear shell modelling of the main walls
to define at which drift the strain limits are reached.
During the nonlinear static analyses, it has been noted that the estimation of the standard engineering quantities
used in international codes (ductility factor Rd and ductility demand μ) suffered great fluctuation. The estima-
tions of such quantities varied depending on the selected analysis method and most of all on the idealization
of the building response. The idealization procedures proposed in the literature, notably from the FEMA-356
prestandard, lead to values not very representative of the state of deformation and solicitation of the reference
building. This raises again questions concerning the adequacy of U.S. procedures when applied directly to the
typical Colombian practices. This also underlines that a thorough understanding of the used analysis methods
and critical interpretation of the results is crucial to avoid mislead use of the results and erroneous conclusion
regarding the seismic behaviour of a building.

72
Colombian RC wall buildings assessment – References

REFERENCES

[1] Asociación Colombiana de Ingeniería Sísmica, NSR-98: Normas Colombianas de Diseño y


Construcción Sismo Resistente, Bogotá DC, Colombia, 1998.
[2] Comisión asesora permanente para el régimen de construcciones sismoresistente, NSR-10: Reglamento
Colombiano de Constructión Sismo Resistente, Bogotá D.C., Colombia: Ministerio de Ambiente,
Vivienda y Desarrollo Territorial, 2010.
[3] American Society of Civil Engineers, ASCE 7-98: Minimum Design Loads for Buildings and Other
Structures, Reston, Virginia, USA, 2008.
[4] American Concrete Institute, ACI 318M-08: Building Code Requirements for Structural Concrete and
Commentary, Farmington Hills, Michigan, USA, June 2008.
[5] American Concrete Institute, ACI 318M-11: Building Code Requirements for Structural Concrete and
Commentary, Farmington Hills, Michigan, USA, 2011.
[6] Swiss Society of Engineers and Architects, SIA 261:2003: Actions on Structures, Zurich, Switzerland:
SN: Schweizer Norm, Norme Suisse, Norma svizzera, 2003.
[7] Deutsches Institut für Normung e.V., DIN 4149: Bauten in deutschen Erdbebengebieten - Lastannahmen,
Bemessung und Ausführung üblicher Hochbauten, Berlin, Germany, 2005.
[8] Comisión asesora permanente para el régimen de construcciones sismoresistentes, NSR-10: Anexo:
Modificaciones técnicas y científicas al Reglamento Colombiano de Construcción Sismo Resistente,
Bogotá D.C., Colombia: Ministerio de Ambiente, Vivienda y Desarrollo Territorial, 2010.
[9] Swiss Society of Engineers and Architects, SIA 260:2003: Basis of Structural Design, Zurich,
Switzerland: SN: Schweizer Norm, Norme Suisse, Norma svizzera, 2004.
[10] European Committee for Standardization, EN 1990: Eurocode 0 - Basis of Structural design, Brussels,
Belgium: European Standard, 2002.
[11] Ministerio de Vivienda, Construcción y Saneamiento, E.60: Norma Técnica de Edificación, Concreto
Armado, Lima, Peru, May 2009.
[12] Comité Ejecutivo de la Norma Ecuatoriana de la Construccion, NEC-11: Norma Ecuatoriana de la
Construcción, Estructuras de Hormigón Armado, Quito, Ecuador, 2011.
[13] Swiss Society of Engineers and Architects, SIA 262:2013: Concrete Structures, Zurich, Switzerland:
SN: Schweizer Norm, Norme Suisse, Norma svizzera, 2013.
[14] European Committee for Standardization, EN 1992: Eurocode 2 - Design of concrete structures,
Brussels, Belgium: European Standard, 2004.
[15] Asociación Colombiana de Ingeniería Sismica, CCC SR-84: Codigo Colombiano de Construcciones
Sismo-Resistentes, Bogotá D.C, Colombia, 1984.
[16] American Society of Civil Engineers, ASCE 031-03: Seismic Evaluation of Existing Buildings, Reston,
Virginia, USA, 2003.
[17] Applied Technology Council, ATC-40: Seismic evaluation and Retrofit of Concrete Buildings, Vol. 1,
Appendices, Vol. 2, Redwood City, California, USA, 1996.

73
Colombian RC wall buildings assessment – References

[18] Federal Emergency Management Agency, Building Seismic Safety Council, FEMA 178: NEHRP
Handbook for Seismic Evaluation of Existing Buildings, Washington D.C., USA, 1992.
[19] Swiss Society of Engineers and Architects, SIA 269:2011: Bases pour la maintenance des structrures
porteuses, Zurich, Switzerland: SN: Schweizer Norm, Norme Suisse, Norma svizzera, 2013.
[20] European Committee for Standardization, EN 1998-1: Eurocode 8 - Design of Structures for Earthquake
Resistance, Part 1: General Rules, Seismic Actions and Rules for Buildings, Brussels, Belgium:
European Standard, 2004.
[21] T. Paulay and M. Priestley, Seismic Design of Reinforced Concrete and Masonry Buildings, New York:
John Wiley & Sons., 1992.
[22] Federal Emergency Management Agency, American Society of Civil Engineers, FEMA-356:
Prestandard and Commentary for the Seismic Rehabilitation of Buildings, Washington DC, USA, 2000.
[23] Federal Emergency Management Agency, SAC Joint Venture, FEMA-350: Recommended Seismic
Design Criteria for New Steel Moment-Frame Buildings, USA, June 2000.
[24] T. Paulay and M. Priestley, “Stability of Ductile Structural Walls,” ACI Structural Journal, pp. 385-392,
July-August 1993.
[25] M. Priestley, G. Calvi and M. Kowalski, Displacement-Based Seismic Design of Structures, Pavia, Italy:
IUSS Press, Istituto Universitario di Studi Superiori di Pavia, 2007.
[26] Grupo de Sismología de Medellín, “Microzonificación Sísmica,” Sistema Municipal Para la Prevención
y Atención de Desastres (SIMPAD), Alcadía de Medellín, Medellín, Colombia, 2011.
[27] F. J. Vecchio and M. P. Collins, “The Modified Compression-Field Theory for Reinforced Concrete
Elements Subjected to Shear,” ACI Journal, pp. 219-231, March-April 1986.
[28] F. J. Vecchio, “Disturbed Stress Field Model for Reinforced Concrete: Formulation,” Journal of
Structural Engineering, vol. January, pp. 12-20, 2001.
[29] SeismoSoft, “SeismoStruct v6.5,” Pavia, Italy, 2013.
[30] J. Mander, M. Priestley and R. Park, “Theoretical Stress-Strain Model for Confined Concrete,” Journal
of Structural Engineering, vol. 114, no. 8, pp. 1804-1826, 114(8) 1988.
[31] J. Martinez-Rueda and A. Elnashai, “Confined Concrete Model unde Cyclic Load,” Materials and
Structures/Matériaux et Constructions, vol. 30, pp. 139-147, April 1997.
[32] M. Menegotto and P. Pinto, “Method of Analysis for Cyclically Loaded R.C. Plane Frames Including
Changes in Geometry and Non-Elastic Behaviour of Elements under Combined Normal Force and
Bending,” in Symposium on the Resistance and Ultimate Deformability of Structures Acted on by Well
Defined Repeated Loads, International Association for Bridge and Structural Engineering, Reports of
the Working Commissions, Zurich, Switzerland, 1973, pp. 15-22.
[33] F. Filippou, E. Popov and V. Bertero, Effects of Bond Deterioration on Hysteretic Behavior of
Reinforced Concrete Joints, Berkeley, California, USA, August 1983.
[34] European Committee for Standardization, EN 1998-3: Eurocode 8 - Design of Structures for Earthquake
Resistance, Part 3: Assessment and Retrofitting of Buildings, Brussels, Belgium: European Standard,
2005.
[35] European Committee for Standardization, EN 1998-3 Corrigendum: Eurocode 8 - Design of Structures
for Earthquake resistance - Part 3: Assessment and Retrofitting of Buildings - Corrigendum, Brussels,
Belgium: European Standard, 2013.
[36] K. P. Abani, “School of engineering and Applied Science of the University of Buffalo, the State
University of New York,” [Online]. Available: http://www.eng.buffalo.edu/~abani/fem/dyn/dyn.html.
[Accessed 06 08 2014].

74
Colombian RC wall buildings assessment – References

[37] P. Fajfar, “Capacity Spectrum Method Based on Inelastic Demand Spectra,” Earthquake Engineering
and Structural Dynamics, no. 28, pp. 979-993, 1999.

75
Colombian RC wall buildings assessment – Appendix A: List of model files

APPENDIX A: LIST OF MODEL FILES

The present appendix lists the model files and postprocessing files used to plot each graph present in the the-
sis, as well as the direction of the folder they are located in.
Figure 3.5 – Elastic displacement response spectra
 Model files:
o None
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\Reference_Building_SS\ 6-Verification demand Monte Robles.xlsx
Figure 4.4 – Stress-strain curves of the tensile test on the Ø6 and Ø16 rebars
 Model files:
o None
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\TW1\ TU1_TensileTestD6&D16_22072014.xlsx
Figure 4.8 – Stress-strain curves for steel rebars for VecTor2 model and selected experimental
results
 Model files:
o None
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\TW1\ TU1_TensileTestD6&D16_22072014.xlsx
Figure 4.10 – Stress-strain relationship of the concrete in the SeismoStruct model
 Model files:
o Files_Thesis_MC\TW1\SeismoStruct\TW1_A_v004.spf
 Matlab and/or excel postprocessing files:
o None
Figure 4.12 – Stress-strain evolution for the concrete of the bottom 4 elements of the free ex-
tremity of the web in the VecTor2 analysis when pushing in the positive direction
 Model files:
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\CO_10m_5%_A_adapted.fwx
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Read_geometry.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Post_processing.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Strain_displ_response_strain_in_concrete.m
Figure 4.13 – Stress-strain relationship of the rebars in the SeismoStruct model (a) Ø6 (b) Ø16
 Model files:
o Files_Thesis_MC\TW1\SeismoStruct\TW1_A_v004.spf
 Matlab and/or excel postprocessing files:
o None
Figure 4.14 – Base shear vs top displacement: Comparison between numerical models (VecTor2
and Seismostruct) and experimental results
 Model files:
o Files_Thesis_MC\TW1\SeismoStruct\TW1_A_v004.spf

A.1
Colombian RC wall buildings assessment – Appendix A: List of model files

o Files_Thesis_MC\TW1\SeismoStruct\TW1_B_v004.spf
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\CO_10m_5%_A_adapted.fwx
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\CO_10m_5%_B_adapted.fwx
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Read_geometry.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Post_processing.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Force_displ_response.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Read_geometry.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Post_processing.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Force_displ_response.m
o Files_Thesis_MC\TW1\TW1_comparisons_v004.xslx
Figure 4.15 – Vertical stress at the base vs top horizontal displacement comparison, flange side
 Model files:
o Files_Thesis_MC\TW1\SeismoStruct\TW1_A_v004.spf
o Files_Thesis_MC\TW1\SeismoStruct\TW1_B_v004.spf
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\CO_10m_5%_A_adapted.fwx
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\CO_10m_5%_B_adapted.fwx
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Read_geometry.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Post_processing.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Strains.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Strain_displ_response.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Read_geometry.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Post_processing.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Strains.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Strain_displ_response.m
o Files_Thesis_MC\TW1\TW1_comparisons_v004.xslx
Figure 4.16 – Vertical stress at the base vs top horizontal displacement comparison, web side
 Model files:
o Files_Thesis_MC\TW1\SeismoStruct\TW1_A_v004.spf
o Files_Thesis_MC\TW1\SeismoStruct\TW1_B_v004.spf
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\CO_10m_5%_A_adapted.fwx
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\CO_10m_5%_B_adapted.fwx
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Read_geometry.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Post_processing.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Strains.m

A.2
Colombian RC wall buildings assessment – Appendix A: List of model files

o Files_Thesis_MC\TW1\VecTor2\10m_5%_A_Adapted\01_Matlab_Postpro-
cessing\Strain_displ_response.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Read_geometry.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Post_processing.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Strains.m
o Files_Thesis_MC\TW1\VecTor2\10m_5%_B_Adapted\01_Matlab_Postpro-
cessing\Strain_displ_response.m
o Files_Thesis_MC\TW1\TW1_comparisons_v004.xslx
Figure 5.4 – Deformed shapes of the 5 first vibration modes of the reference building model
 Model files:
o Files_Thesis_MC\Reference_Building_SS\Eigenvalues\ V024_coef115_Rzblocked_eigen-
values.spf
 Matlab and/or excel postprocessing files:
o None
Figure 5.6 – Global response of the reference building
 Model files:
o Files_Thesis_MC\Reference_Building_SS\X-\V024_coef115_Rzblocked_X-.spf
o Files_Thesis_MC\Reference_Building_SS\X+\V024_coef115_Rzblocked_X+.spf
o Files_Thesis_MC\Reference_Building_SS\Y-\V024_coef115_Rzblocked_Y-.spf
o Files_Thesis_MC\Reference_Building_SS\Y+\V024_coef115_Rzblocked_Y+.spf
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\Reference_Building_SS\2_StrainHardenindInIdealization_GLOBAL.xlsx
Figure 5.7 – EPSH1 and EPSH2 idealizations of the building response in direction X+
 Model files:
o Files_Thesis_MC\Reference_Building_SS\X+\V024_coef115_Rzblocked_X+.spf
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\Reference_Building_SS\2_StrainHardenindInIdealization_GLOBAL.xlsx
o Files_Thesis_MC\Reference_Building_SS\2_StrainHardenindInIdealiza-
tion_GLOBAL_handidealization.xlsx
Figure 5.8 – EPSH1 and EPSH2 idealizations of the building response in direction Y+
 Model files:
o Files_Thesis_MC\Reference_Building_SS\Y+\V024_coef115_Rzblocked_Y+.spf
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\Reference_Building_SS\2_StrainHardenindInIdealization_GLOBAL.xlsx
o Files_Thesis_MC\Reference_Building_SS\2_StrainHardenindInIdealiza-
tion_GLOBAL_handidealization.xlsx
Figure 5.9 – EPP1 and EPP2 idealizations of the building response in direction X+
 Model files:
o Files_Thesis_MC\Reference_Building_SS\X+\V024_coef115_Rzblocked_X+.spf
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\Reference_Building_SS\2_ElasticPerfectlyPlasticIdealiza-
tion_GLOBAL.xlsx
o Files_Thesis_MC\Reference_Building_SS\2_ElasticPerfectlyPlasticIdealiza-
tion_GLOBAL_handidealization.xlsx
Figure 5.10 – EPP1 and EPP2 idealizations of the building response in direction Y+
 Model files:
o Files_Thesis_MC\Reference_Building_SS\X+\V024_coef115_Rzblocked_Y+.spf

A.3
Colombian RC wall buildings assessment – Appendix A: List of model files

 Matlab and/or excel postprocessing files:


o Files_Thesis_MC\Reference_Building_SS\2_ElasticPerfectlyPlasticIdealiza-
tion_GLOBAL.xlsx
o Files_Thesis_MC\Reference_Building_SS\2_ElasticPerfectlyPlasticIdealiza-
tion_GLOBAL_handidealization.xlsx
Figure 5.11 – Acceleration-Displacement Response Spectrum of the equivalent SDOF in direc-
tion X+
 Model files:
o Files_Thesis_MC\Reference_Building_SS\X+\V024_coef115_Rzblocked_X+.spf
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\Reference_Building_SS\2_ElasticPerfectlyPlasticIdealiza-
tion_GLOBAL.xlsx
o Files_Thesis_MC\Reference_Building_SS\2_ElasticPerfectlyPlasticIdealiza-
tion_GLOBAL_handidealization.xlsx
Figure 5.12 – Acceleration-Displacement Response Spectrum of the equivalent SDOF in direc-
tion Y+
 Model files:
o Files_Thesis_MC\Reference_Building_SS\X+\V024_coef115_Rzblocked_Y+.spf
 Matlab and/or excel postprocessing files:
o Files_Thesis_MC\Reference_Building_SS\3_ElasticPerfectlyPlasticIdealization_SDOF.xlsx
o Files_Thesis_MC\Reference_Building_SS\3_ElasticPerfectlyPlasticIdealiza-
tion_SDOF_handidealization.xlsx

A.4

Potrebbero piacerti anche