seismic assessment of reinforced concrete frame structures

477
Seismic Assessment of Reinforced Concrete Frame Structures with a New Flexibility Based Element António José Coelho Dias Arêde Thesis submitted to the Faculdade de Engenharia da Universidade do Porto in candidature for the degree of Doutor in Civil Engineering FACULDADE DE ENGENHARIA UNIVERSIDADE DO PORTO October 1997 Publication financially supported by the Junta Nacional de Investigação Científica e Tecnológica

Upload: others

Post on 17-Jun-2022

11 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Seismic Assessment of Reinforced Concrete Frame Structures

Seismic Assessment of Reinforced Concrete Frame Structures with a New Flexibility Based Element

António José Coelho Dias Arêde

Thesis submitted to the Faculdade de Engenharia da Universidade do Porto in candidature for the degree of Doutor in

Civil Engineering

FACULDADE DE ENGENHARIA

UNIVERSIDADE DO PORTO

October 1997

Publication financially supported by the Junta Nacional de Investigação Científica e Tecnológica

Page 2: Seismic Assessment of Reinforced Concrete Frame Structures
Page 3: Seismic Assessment of Reinforced Concrete Frame Structures
Page 4: Seismic Assessment of Reinforced Concrete Frame Structures

Il faut avoir le courage de dire des

choses imparfaites, de renoncer au

mérite d’avoir fait tout ce qu’on pouvait

faire, d’avoir dit tout ce qu’on pouvait

dire, enfin de sacrifier son amour-propre

au désir d’être utile et d’améliorer la

marche du progrès.

Lavoisier

Page 5: Seismic Assessment of Reinforced Concrete Frame Structures

ACKNOWLEDGEMENTS

I wish to express my sincere gratitude to the following persons and institutions, which have

contributed to the accomplishment of this milestone of my academic education and profes-

sional career:

• To Professor Raimundo Delgado, my supervisor, whose collaboration and friendship went

far beyond the academic and technical field. His comments and suggestions have always

played a very important role in the course of this study.

• To Artur Pinto, for the opportunity of staying at the ELSA laboratory of the Joint Research

Centre at Ispra, for his extremely valuable support and patient guidance, and, very specially,

for the friendship and confidence. Last but not the least, I thank him and his family very

much for the solidarity in many circumstances, which has definitely contributed for my

social and human integration in Italy.

• To João Guedes, my housemate, for the friendship, understanding and tolerance shown dur-

ing three and half years of social and professional life. Also, to Alfredo Campos Costa, for

the precious help, for the fruitful discussions and the encouraging words in some difficult

situations.

• To Pierre Pegon, for the excellent suggestions and advice in the model development and

implementation.

• To Professor Couto Marques for the precision and remarkable patience in revising the text.

• To all ELSA staff for the friendly and welcoming environment, with particular reference to

Professor Jean Donea for the confidence and the establishment of further cooperation links.

• To my friends and colleagues Rui Faria and Nelson Vila Pouca, for having so promptly

released me from my teaching activity, which was determinant for my intense dedication to

thesis writing.

• To my parents, brothers and nephews for all the care, patience and understanding during all

this long period of absence. To a large extent, I owe them the happy end of this and other

steps of my life.

• To my very good friends Tó Viana and Luísa who have always known how to be supportive

and encouraging in all situations. They are the living proof that the Friendship is not a void

word.

Page 6: Seismic Assessment of Reinforced Concrete Frame Structures

• To all my Ispra friends with whom I have shared excellent moments and who have helped

me feel at home in a foreign country.

Part of the research reported in this thesis was financially supported by the Human Capital and

Mobility programme of the European Commission, under the PREC8 (“Prenormative

Research in Support of Eurocode 8”) project. A one-year grant and the publication of this the-

sis were supported by the Portuguese Board for Scientific and Technological Research (JNICT

- “Junta Nacional de Investigação Científica e Tecnológica”). Both financial supports are

gratefully acknowledged.

Page 7: Seismic Assessment of Reinforced Concrete Frame Structures

Abstract

The present thesis focuses on the development of a global element model for the non-linear analysis ofreinforced concrete (RC) frame structures when subjected to monotonic or cyclic loads. The model isvalidated with the results of a broad experimental campaign on a full scale structure and is intensivelyapplied to the seismic behaviour assessment of structures designed according to Eurocode 8.

An innovative flexibility-based member model is presented, where the flexibility formulation isadopted to avoid the difficulties in accounting for the modifications in the kinematic shape functions ofthe classic stiffness formulation due to progressive variation of the element stiffness during the loadinghistory. The flexibility formulation makes use of force shape functions which are strictly derived fromequilibrium conditions and, thus, remain exact regardless of the element state.

The non-linear behaviour is controlled by a section moment-curvature model of Takeda type with trilin-ear skeleton curves. Besides the member-end sections and one mid-span fixed section, a number ofmoving control sections are monitored in order to constantly define and update the uncracked, crackedand yielded zones inside the element. For a given moment distribution (corresponding to imposed end-section rotations) the location of the moving control sections is first updated. The flexibility and thecurvature distributions along the element are then defined according to the referred section models,leading, by integration, to the element flexibility matrix and the end-section rotations, respectively. Aninternal iterative scheme is required to ensure that the curvature distribution in the element leads to end-section rotations compatible with the imposed ones and, at convergence, both the plastic zone lengthsand the progressive softening due to cracking become automatically defined.

Such a modelling strategy allows the element state along its full length to be updated at each load stepand, consequently, provides an adequate simulation of both the global structural stiffness and the evolu-tion of dynamic characteristics during the seismic response. The model is implemented in a generalpurpose computer code for finite element static and dynamic structural analysis, together with an auxil-iary procedure for the trilinear skeleton curve definition based on an efficient algorithm specificallydesigned to avoid the usual fibre discretization of the section.

The model is used to simulate the seismic response of a four-storey full scale RC frame structurepseudo-dynamically tested for two different earthquake levels and quasi-statically tested with increas-ing intensity cyclic load up to a near-failure stage. The tests, carried out at the ELSA laboratory of theJoint Research Centre at Ispra (Italy), are fully described and the test results are compared against thenumerical simulations. This provides an excellent means of checking whether the model is able todescribe the quasi-static or dynamic structural behaviour throughout distinct stages, while keeping agood compromise between computational efficiency and result quality.

The non-linear seismic analysis of a set of RC frame structures designed according to Eurocode 8 iscarried out in the framework of a European-wide prenormative research programme in support of thatdesign code (PREC8). Structures consist of two basic configurations (one regular and another irregu-lar), designed for several combinations of ductility class and design acceleration. Numerical modellingis done by means of the proposed flexibility element model and the seismic analysis is performed con-sidering several accelerograms fitting the EC8 spectrum and scaled by increasing intensities. The struc-tural responses are analysed by relative comparison between trial cases, focusing on issues such as:overstrength, cracking and yielding patterns, local ductility and damage distribution, drift and damage.An exercise of system reliability analysis is also presented in order to estimate bounds of failure proba-bility of the various structures, which are then compared between trial cases in order to find out theinfluence of the design parameters on the structural safety.

Page 8: Seismic Assessment of Reinforced Concrete Frame Structures

Resumo

Na presente tese procede-se ao desenvolvimento de um modelo de elemento global destinado à análisenão-linear de estruturas em pórtico de betão armado, quando sujeitas a acções monotónicas ou alterna-das. O modelo é validado por meio da simulação de uma vasta campanha de ensaios experimentaissobre uma estrutura testada à escala real, e é aplicado na verificação do comportamento sísmico deestruturas projectadas de acordo com o Eurocódigo 8 (EC8).

Apresenta-se um modelo inovador, desenvolvido com base na formulação de flexibilidade a fim decontornar as dificuldades associadas às modificações das funções de forma cinemáticas da formulaçãoclássica de rigidez, que são originadas pela alteração progressiva da rigidez do elemento durante ahistória de carga. A formulação de flexibilidade baseia-se em funções de forma de força que são obtidasexclusivamente por condições de equilíbrio, pelo que se mantêm exactas independentemente do estadodo elemento.

O comportamento não-linear é controlado por um modelo de secção do tipo Takeda, com curvas basetrilineares em termos de momento-curvatura. Para além das secções de extremidade e de uma secçãocentral fixa, são ainda controladas certas secções móveis que permitem definir e actualizar constante-mente as zonas plastificadas, fendilhadas e não-fendilhadas dentro do elemento. Para uma dada dis-tribuição de momentos (correspondente a rotações impostas nas extremidades), a posição das secçõesmóveis é devidamente actualizada. Subsequentemente, as distribuições de flexibilidade e de curvaturaao longo do elemento são definidas de acordo com os modelos de secção e, por integração, obtêm-se amatriz de flexibilidade e as rotações de extremidade, respectivamente. Um processo iterativo internoencarrega-se de garantir que uma dada distribuição de curvaturas origina rotações de extremidade com-patíveis com as rotações impostas e, atingida a convergência, tanto os comprimentos das zonas plásti-cas como a progressiva perda de rigidez devida à fendilhação vêm automaticamente definidos.

Esta técnica de modelação permite a actualização do estado do elemento em todo o seu comprimento acada passo de carga, conduzindo assim a uma simulação adequada da rigidez global da estrutura e daevolução das suas características dinâmicas durante a resposta sísmica. O modelo é incorporado numprograma geral de análise estrutural estática e dinâmica por elementos finitos, conjuntamente com umprocedimento numérico auxiliar para a definição da curva base trilinear baseado num algoritmo eficazespecificamente desenvolvido para evitar a discretização da secção em fibras.

O modelo de elemento é usado para reproduzir numericamente a resposta dum modelo físico à escalareal duma estrutura em pórtico de betão armado com quatro pisos, testado com recurso ao métodopseudo-dinâmico para duas intensidades sísmicas diferentes, e também sob acções cíclicas de intensid-ades gradualmente crescentes até um estado próximo da ruína. Os testes, realizados no laboratórioELSA do centro Comum de Investigação em Ispra (Itália), são detalhadamente descritos e os resultadossão comparados com os das simulações numéricas. Tal comparação fornece num excelente meio de val-idação que permite verificar se o modelo numérico é adequado para descrever o comportamento estru-tural quase-estático ou dinâmico ao longo de diversas fases, garantindo um bom compromisso entreeficiência computacional e qualidade de resultados.

Apresenta-se a análise sísmica não-linear de algumas estruturas em pórtico de betão armado projecta-das segundo o EC8, no quadro de um programa europeu de investigação de validação deste eurocódigo(PREC8). As estruturas consistem em duas configurações básicas, uma regular e outra irregular, pro-jectadas para várias classes de ductilidade e acelerações de projecto. A modelação numérica é feitacom o elemento global proposto e a análise sísmica é realizada com vários acelerogramas, gerados porforma a ajustarem-se ao espectro de projecto do EC8, e escalados com intensidades crescentes. As res-postas estruturais são analisadas por comparação relativa entre os diversos casos, concentrando especi-ficamente a atenção em aspectos como: a sobre-resistência, distribuições de fendilhação, plastificação,exigência de ductilidade e dano local, "drift" e dano global. É incluído também um exercício de cálculode limites das probabilidades de ruína, que são comparados entre as várias estruturas a fim de averiguara influência de alguns parâmetros de projecto na segurança estrutural.

Page 9: Seismic Assessment of Reinforced Concrete Frame Structures

Resumé

Cette thèse se consacre au développement d’un modèle d’élément global pour l’analyse non-linéairedes structures formées de portiques en béton armé sous chargements monotones ou cycliques. Lemodèle est d’abord validé à l’aide des résultats d’un large ensemble d’essais expérimentaux effectuéessur une structure à l’échelle réelle, puis appliqué massivement à la vérification du comportement sis-mique des structures dimensionnées avec l’Eurocode 8 (EC8).

Un nouveau modèle d’élément est présenté, dont la formulation de flexibilité a été choisie pour éviterles difficultés associées aux modifications des fonctions de forme cinématiques dans la formulationclassique de rigidité, modifications qui résultent de la variation progressive de la raideur de l’élémentpendant la réponse. La formulation de flexibilité utilise des fonctions de forme des forces, strictementobtenues à partir des conditions d’équilibre, et donc exactes indépendamment de l’état de l’élément.

Le comportement non-linéaire est controlé par des lois moment-courbure du type Takeda, basées surdes courbes enveloppes trilinéaires. Outre les deux sections d’extrémité et une section centrale fixe,quelques sections mobiles sont aussi contrôlées pour définir et actualiser constamment les zones plasti-fiées, fissurées et non-fissurées dans l’élément. Pour une certaine distribution de moments fléchissants(correspondant à des rotations imposées dans les noeuds), la position des sections mobiles est d’abordactualisée. Puis, les distributions de flexibilité et courbure dans l’élément sont définies à l’aide desmodèles de section. Finalement, la matrice de flexibilité et les rotations aux extrémités peuvent êtredéduites par intégration. Des itérations internes sont nécessaires pour assurer que la distribution decourbure soit compatible avec les rotations imposées aux extrémités; lors que la convergence estatteinte, les zones plastiques et l’adoucissement progressif dû à la fissuration sont automatiquementdéfinis.

Ce type de modélisation permet l’actualisation de l’état de l’élément dans toute sa longueur à chaquepas de charge, ce qui fournit une simulation convenable de la raideur globale de la structure ansi que del’évolution de ses caractéristiques dynamiques pendant la réponse sismique. Le modèle a été introduitdans un code de calcul général d’éléments finis pour l’analyse non-linéaire, statique ou dynamique, destructures, et une procédure auxiliaire a été écrite pour la définition des courbes enveloppes trilinéairesbasée sur un algoritme spécifique afin d’éviter des discrétisations à fibres.

Le modèle d’élément est utilisé pour la simulation numérique de la réponse sismique d’une structure àquatre étages formée de portiques en béton armé, testée avec la méthode pseudo-dynamique sousl’action de séismes de deux intensités différentes, ainsi que sous chargement cyclique quasi-statiquejusqu’à un état proche de la ruine. Les tests, qui ont été faits au laboratoire ELSA du Centre Communde Recherche à Ispra (Italie), sont complètement décrits et les résultats sont comparés avec ceux dessimulations numériques. Cela constitue un excellent moyen de vérifier si le modèle peut reproduire laréponse quasi-statique ou dynamique dans diverses phases de comportement, en conservant un bon rap-port efficace computationelle versus qualité des résultats.

L’analyse sismique non-linéaire de quelques structures en béton armé, conçues et dimensionées selonl’EC8, est faite dans le cadre d’un programme européen de recherche prénormative (PREC8). Deuxconfigurations structurelles sont considérées (une regulière et une autre irregulière), dimensionnéespour différentes combinaisons de classe de ductilité et d’accélération de projet. La modélisationnumérique est faite avec l’élément développé et quelques accélérogrammes compatibles avec le spectrede l’EC8 sont utilisés (aux intensités croissantes) pour l’analyse sismique. Les réponses structurellessont comparées entre les différents cas, en ce qui concerne la réserve de résistance, les distributions defissuration, plastification, ductilité et endommagement local, ainsi que les "drifts" et l’endommagementglobal. Un exercice de calcul des limites de la probabilité de ruine est aussi fait, pour vérifier l’influ-ence de quelques paramètres de projet sur la sécurité structurelle.

Page 10: Seismic Assessment of Reinforced Concrete Frame Structures

TABLE OF CONTENTS

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxi

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxix

Chapter 1

INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51.3 Chapter organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

Chapter 2

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART . . 13

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132.2 Hysteretic behaviour of reinforced concrete members . . . . . . . . . . . . . . . . . . . . . . . . 142.3 Numerical modelling strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.3.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192.3.2 Member type models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212.3.3 Distributed inelasticity member models . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.4 General flexibility formulation for beam-column elements . . . . . . . . . . . . . . . . . . . . . . 332.4.1 Conventions and notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332.4.2 Relations between spaces of variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372.4.3 Theoretical background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

2.4.3.1 Stiffness method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402.4.3.2 Flexibility method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

2.4.4 The element state determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452.4.4.1 General comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452.4.4.2 Nodal forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472.4.4.3 Element loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502.4.4.4 Remarks on the global non-linear algorithm . . . . . . . . . . . . . . . . . . . . 592.4.4.5 Control sections and numerical integration . . . . . . . . . . . . . . . . . . . . 61

2.5 Concluding summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

Chapter 3

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL. . . . . . 63

3.1 General comments and innovative features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633.2 Basic assumptions and remarks on convention and notation . . . . . . . . . . . . . . . . . . . . . 66

Page 11: Seismic Assessment of Reinforced Concrete Frame Structures

xvi TABLE OF CONTENTS

3.3 Trilinear model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703.4 Control sections and element zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.4.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723.4.2 Cracking sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 743.4.3 Yielding sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 823.4.4 Null moment sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

3.5 Behaviour of the control sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 853.5.1 Modified trilinear model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

3.5.1.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 853.5.1.2 Motivation for model modification . . . . . . . . . . . . . . . . . . . . . . . . . 863.5.1.3 Proposed model modifications . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

3.5.2 Transition from uncracked to cracked section behaviour . . . . . . . . . . . . . . . . . . . 933.5.3 State evolution of control sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

3.5.3.1 Internal moving sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 973.5.3.2 Fixed sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

3.6 Element state determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1003.6.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1003.6.2 Flexibility distribution within the element . . . . . . . . . . . . . . . . . . . . . . . . . 1013.6.3 Element flexibility matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1033.6.4 Displacement residuals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1083.6.5 Element applied loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1113.6.6 Behaviour of plastic end zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

3.6.6.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1123.6.6.2 Plastic zone splitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1153.6.6.3 Event-to-event scheme in the element iterative process . . . . . . . . . . . . . 1183.6.6.4 Evolution of curvatures in fixed plastic zones . . . . . . . . . . . . . . . . . . 120

3.6.7 Integration of deformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1223.6.8 Convergence criteria for the element iterative process . . . . . . . . . . . . . . . . . . . 1243.6.9 Convergence problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

3.6.9.1 Difficult or no- convergence situations . . . . . . . . . . . . . . . . . . . . . . 1253.6.9.2 Line search scheme for element iterations . . . . . . . . . . . . . . . . . . . . 128

3.7 Summary of the non-linear algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1313.7.1 General structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1313.7.2 Element state determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

3.8 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

Chapter 4

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION . . . . . . . . . . . 141

4.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1414.2 Implementation in the computer code CASTEM2000 . . . . . . . . . . . . . . . . . . . . . . . . 142

4.2.1 Basics of CASTEM2000 and main implementation needs . . . . . . . . . . . . . . . . . 1424.2.2 Flexibility based element implementations . . . . . . . . . . . . . . . . . . . . . . . . . 1494.2.3 Definition of skeleton curves for RC global section modelling . . . . . . . . . . . . . . 153

4.2.3.1 Type of sections, notations and conventions . . . . . . . . . . . . . . . . . . . 1534.2.3.2 Material models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1554.2.3.3 Linear behaviour: the cracking point . . . . . . . . . . . . . . . . . . . . . . . 158

Page 12: Seismic Assessment of Reinforced Concrete Frame Structures

TABLE OF CONTENTS xvii

4.2.3.4 Turning points for non-linear behaviour . . . . . . . . . . . . . . . . . . . . . 1604.2.3.5 Remarks on implementation and validation . . . . . . . . . . . . . . . . . . . 175

4.3 Flexibility-based element validation at the single member level . . . . . . . . . . . . . . . . . . 1764.3.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1764.3.2 Specimen characteristics and test description . . . . . . . . . . . . . . . . . . . . . . . 1764.3.3 Numerical simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1804.3.4 Remarks on model validation results . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

4.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

Chapter 5

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA . . . . . . . . . . . . . . . . . . . . 191

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1915.2 The Pseudo-Dynamic test method. An overview . . . . . . . . . . . . . . . . . . . . . . . . . . 192

5.2.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1925.2.2 Time integration techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193

5.2.2.1 Newmark explicit algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 1955.2.2.2 α-implicit algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1965.2.2.3 Mixed explicit-implicit algorithms . . . . . . . . . . . . . . . . . . . . . . . . 199

5.2.3 Substructuring in the PSD method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2005.2.4 Applications of the PSD method at the ELSA laboratory . . . . . . . . . . . . . . . . . 202

5.3 Structure design and layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2035.4 Material properties and reduced scale member tests . . . . . . . . . . . . . . . . . . . . . . . . . 2055.5 Full-scale tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

5.5.1 PSD test setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2085.5.2 The input accelerogram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2115.5.3 Preliminary tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2135.5.4 Seismic tests on the bare frame structure . . . . . . . . . . . . . . . . . . . . . . . . . 214

5.5.4.1 Low level test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2145.5.4.2 High level test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219

5.5.5 Infilled frame seismic tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2215.5.5.1 Uniformly infilled configuration . . . . . . . . . . . . . . . . . . . . . . . . . 2225.5.5.2 Soft-storey configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225

5.5.6 Final cyclic tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2265.6 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

Chapter 6

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING. . . . . . . . . . . . . . . . . . . . . . . 243

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2436.2 Modelling assumptions and data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244

6.2.1 Structure modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2446.2.1.1 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2446.2.1.2 Collaborating slab width . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2456.2.1.3 Mass and vertical static loads . . . . . . . . . . . . . . . . . . . . . . . . . . 248

6.2.2 Cross-section characteristics and material properties . . . . . . . . . . . . . . . . . . . 249

Page 13: Seismic Assessment of Reinforced Concrete Frame Structures

xviii TABLE OF CONTENTS

6.2.2.1 Cross-sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2496.2.2.2 Concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2526.2.2.3 Steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253

6.2.3 Skeleton curves for the section model . . . . . . . . . . . . . . . . . . . . . . . . . . . 2536.2.4 Hysteretic behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255

6.2.4.1 Unloading stiffness degradation . . . . . . . . . . . . . . . . . . . . . . . . . 2556.2.4.2 Pinching effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2566.2.4.3 Strength degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256

6.2.5 Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2576.2.6 Modelling of infills . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258

6.2.6.1 Setup of numerical tools for infill modelling . . . . . . . . . . . . . . . . . . . 2596.2.6.2 Application to a single frame . . . . . . . . . . . . . . . . . . . . . . . . . . . 2616.2.6.3 Application to the full-scale structure . . . . . . . . . . . . . . . . . . . . . . 2626.2.6.4 Infill panels in the present study . . . . . . . . . . . . . . . . . . . . . . . . . 263

6.3 Damage quantification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2646.3.1 General overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2646.3.2 The Park and Ang damage index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269

6.3.2.1 The damage parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2696.3.2.2 Yielding rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2716.3.2.3 Ultimate rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2736.3.2.4 Hysteretic dissipated energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 276

6.4 Analysis of results from numerical simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . 2796.4.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2796.4.2 Procedure for static analytical simulation of the tests . . . . . . . . . . . . . . . . . . . 2806.4.3 Static analysis by flexibility modelling versus experimental tests . . . . . . . . . . . . . 284

6.4.3.1 Pushover analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2846.4.3.2 Bare frame seismic tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2866.4.3.3 Infilled frame tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2976.4.3.4 Final cyclic test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3016.4.3.5 Summary of static analysis results . . . . . . . . . . . . . . . . . . . . . . . . 308

6.4.4 Dynamic analysis by flexibility modelling versus experimental tests . . . . . . . . . . . 3096.4.4.1 Comparison of structural frequencies . . . . . . . . . . . . . . . . . . . . . . . 3106.4.4.2 Low-level test on the bare structure . . . . . . . . . . . . . . . . . . . . . . . 3116.4.4.3 High level test on the bare structure . . . . . . . . . . . . . . . . . . . . . . . 3166.4.4.4 Remarks on energy comparison . . . . . . . . . . . . . . . . . . . . . . . . . . 3216.4.4.5 Summary of dynamic analysis results . . . . . . . . . . . . . . . . . . . . . . 325

6.4.5 Flexibility element versus fixed length plastic hinge (F.H.) modelling . . . . . . . . . . 3266.4.5.1 Assumptions for F.H. modelling . . . . . . . . . . . . . . . . . . . . . . . . . 3266.4.5.2 Discussion on F.H. modelling and comparison with flexibility analysis results . 3286.4.5.3 Summary of F.H. and flexibility modelling comparison . . . . . . . . . . . . . 335

6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 336

Chapter 7

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 . . . 341

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3417.2 The PREC8 project . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342

Page 14: Seismic Assessment of Reinforced Concrete Frame Structures

TABLE OF CONTENTS xix

7.2.1 Basics of EC8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3427.2.2 The RC frame structure topic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344

7.3 The building configurations 2 and 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3467.3.1 General comments and structure layout . . . . . . . . . . . . . . . . . . . . . . . . . . 3467.3.2 Vertical static loads and seismic action . . . . . . . . . . . . . . . . . . . . . . . . . . 3497.3.3 Structure modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351

7.3.3.1 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3517.3.3.2 Mass, damping and natural frequencies . . . . . . . . . . . . . . . . . . . . . 3527.3.3.3 Moment-curvature constitutive relations for global section behaviour . . . . . 354

7.4 Non-linear seismic analysis of building configurations 2 and 6 . . . . . . . . . . . . . . . . . . . 3597.4.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3597.4.2 Structural strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3597.4.3 Cracking, yielding and damage patterns . . . . . . . . . . . . . . . . . . . . . . . . . . 3647.4.4 Ductility demand and damage distribution in elevation . . . . . . . . . . . . . . . . . . 3707.4.5 Overall analysis of response parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 3767.4.6 Safety assessment by probabilities of failure. An exercise . . . . . . . . . . . . . . . . 383

7.4.6.1 Methodology and assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . 3847.4.6.2 Comparative analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391

7.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393

Chapter 8

FINAL REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397

8.1 Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3978.2 Future developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 404

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407

Appendix A

Linear Elastic Timoshenko Beam Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421

A.1 Section formulation and constitutive relationship . . . . . . . . . . . . . . . . . . . . . . . . . . 421A.2 Element flexibility matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 422A.3 Element displacements as integrated deformations . . . . . . . . . . . . . . . . . . . . . . . . . 426

Appendix B

Trilinear Model Details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431

Appendix C

Internal Force Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437

C.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437C.2 Element applied forces in three directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437C.3 Element applied forces only in the non-linear bending plane . . . . . . . . . . . . . . . . . . . . 439C.4 Moving section abscissas and respective derivatives . . . . . . . . . . . . . . . . . . . . . . . . 441

Page 15: Seismic Assessment of Reinforced Concrete Frame Structures

xx TABLE OF CONTENTS

Appendix D

The Event-to-Event Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445

Appendix E

Non-Linear Dynamic and Static Analysis Scheme. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 447

Page 16: Seismic Assessment of Reinforced Concrete Frame Structures

LIST OF FIGURES

Chapter 2

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART . . 13

Figure 2.1 Typical global response diagrams of beams and columns for monotonic loading . . . . . . . . 16

Figure 2.2 Cyclic global response examples (Carvalho (1993)) . . . . . . . . . . . . . . . . . . . . . . . 17

Figure 2.3 Member modelling: a) two component model (Clough et al. (1965)) and b) one component model (Giberson (1967)) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

Figure 2.4 Spaces of variables and axis systems: from the global to the local element level. . . . . . . . . 34

Figure 2.5 Space of variables at the local section level.. . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

Figure 2.6 Element applied loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

Figure 2.7 Main tasks of the classical stiffness based state determination process . . . . . . . . . . . . . . 42

Figure 2.8 State determination of flexibility based elements: from the global to the local level . . . . . . . 46

Figure 2.9 Flowchart for the element state determination of flexibility based elements . . . . . . . . . . . 47

Figure 2.10 Details of element and section state determination for flexibility based elements . . . . . . . . 49

Figure 2.11 Details of element and section state determination for flexibility based elements with the application of element loads. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

Figure 2.12 Sequence for the application of incremental element and nodal loads . . . . . . . . . . . . . . 54

Figure 2.13 Details of the element state determination for first internal iteration of the first N-R iteration, in the presence of element loads. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

Figure 2.14 Details of the element state determination for the (j>1) internal iterations of the first N-R iteration, in the presence of element loads. . . . . . . . . . . . . . . . . . . . . . . . . . 56

Figure 2.15 State determination of flexibility based elements for displacements corrections relative to the step beginning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

Chapter 3

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL. . . . . . 63

Figure 3.1 Adjustment of local section axis system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

Figure 3.2 Primary or skeleton trilinear curve for the global section model. . . . . . . . . . . . . . . . . . 71

Figure 3.3 Element control sections: fixed sections (E1, E2 and H) and moving section (M) . . . . . . . . 73

Figure 3.4 Distinction and evolution between cracking and cracked sections . . . . . . . . . . . . . . . . 75

Page 17: Seismic Assessment of Reinforced Concrete Frame Structures

xxii LIST OF FIGURES

Figure 3.5 General layout of assumed cracking sections and local abscissas for no element loads or concentrated force applied in H . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

Figure 3.6 Cracking sections and local abscissas for parabolic moment distribution.. . . . . . . . . . . . 77

Figure 3.7 Examples of restricted cracking sections in the presence of distributed force . . . . . . . . . . 79

Figure 3.8 Definition of cracking moment directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Figure 3.9 General layout of assumed yielding and cracking sections . . . . . . . . . . . . . . . . . . . 83

Figure 3.10 Locations of null moment sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

Figure 3.11 Evolution of cracking and yielding section points in the model diagram . . . . . . . . . . . . 85

Figure 3.12 Comparison of fibre and trilinear section modelling formulations. Section, member and model data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

Figure 3.13 Comparison of fibre and trilinear section modelling formulations. Local and global response for two axial load levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

Figure 3.14 Effects of tension-softening in the fibre formulation. . . . . . . . . . . . . . . . . . . . . . . 90

Figure 3.15 Hysteretic rules for the modified trilinear model. . . . . . . . . . . . . . . . . . . . . . . . . 92

Figure 3.16 Cracking transition in the case of non-uniform moment distribution . . . . . . . . . . . . . . 94

Figure 3.17 Cracking transition in the case of uniform moment distribution . . . . . . . . . . . . . . . . . 95

Figure 3.18 Rules for progressive transition of the cracking plateau transition . . . . . . . . . . . . . . . 96

Figure 3.19 Examples of flexibility distributions for loading and unloading cases. . . . . . . . . . . . . . 102

Figure 3.20 Derivation of additional flexibility terms due to moving sections . . . . . . . . . . . . . . . . 106

Figure 3.21 Monotonic development of plastic zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

Figure 3.22 Plastic zone splitting in fixed and variable length parts . . . . . . . . . . . . . . . . . . . . . 115

Figure 3.23 Flexibility distributions in plastic zones with no further yielding development . . . . . . . . . 116

Figure 3.24 Flexibility distributions in plastic zones with further yielding development . . . . . . . . . . 117

Figure 3.25 Application of the event-to-event scheme to the element state determination . . . . . . . . . . 119

Figure 3.26 Total curvature evolution for non-monotonic loading . . . . . . . . . . . . . . . . . . . . . . 121

Figure 3.27 Cracking transition and the role of additional flexibility terms . . . . . . . . . . . . . . . . . 126

Figure 3.28 Typical cases generating convergence problems. . . . . . . . . . . . . . . . . . . . . . . . . 128

Figure 3.29 Interpolation for line search scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

Figure 3.30 Flow chart for structure state determination . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

Page 18: Seismic Assessment of Reinforced Concrete Frame Structures

LIST OF FIGURES xxiii

Chapter 4

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION . . . . . . . . . . 141

Figure 4.1 Illustrative example for GIBIANE input . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

Figure 4.2 Typical moment-curvature diagram and trilinear approximation . . . . . . . . . . . . . . . . 153

Figure 4.3 Types of sections. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

Figure 4.4 Reference axis system convention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

Figure 4.5 Steel stress-strain relationship . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

Figure 4.6 Concrete stress-strain relationship . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

Figure 4.7 Imminent cracking condition. Section, applied forces and internal strains . . . . . . . . . . . 159

Figure 4.8 Generic situation for the unified procedure. Forces and internal strains . . . . . . . . . . . . 161

Figure 4.9 Rectangular section split into unconfined and confined zones . . . . . . . . . . . . . . . . . 163

Figure 4.10 T-shape section split into unconfined and confined zones . . . . . . . . . . . . . . . . . . . 163

Figure 4.11 Influence of d0 in the development of stress zones . . . . . . . . . . . . . . . . . . . . . . . 164

Figure 4.12 Influence of d0 in the development of uniform geometry zones . . . . . . . . . . . . . . . . 164

Figure 4.13 Development of integration zones and adjustable widths bi and bi . . . . . . . . . . . . . . . 165

Figure 4.14 Yielding criteria at the tensioned steel or at compressed confined concrete . . . . . . . . . . 172

Figure 4.15 Ultimate criteria at the tensioned steel or at compressed confined concrete . . . . . . . . . . 173

Figure 4.16 Supplementary point. Criterion related to the most tensioned steel . . . . . . . . . . . . . . . 174

Figure 4.17 Section layout S2 and schematic representation of tested cantilever beams . . . . . . . . . . 177

Figure 4.18 Cyclic sequences (V5 and V6) of imposed displacements for LNEC and KT beams. . . . . . 179

Figure 4.19 Comparison of numerical and experimental monotonic M-ϕ curves for LNEC and KT beam sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

Figure 4.20 LNEC and KT beams S2: monotonic tests V1 and V2 . . . . . . . . . . . . . . . . . . . . . 183

Figure 4.21 LNEC beam S2: tests V5 and V6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

Figure 4.22 KT beam S2: tests V5 and V6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

Chapter 5

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA . . . . . . . . . . . . . . . . . . . . 191

Figure 5.1 Implementation of the explicit PSD method. . . . . . . . . . . . . . . . . . . . . . . . . . . 196

Figure 5.2 Implementation of the implicit PSD method. . . . . . . . . . . . . . . . . . . . . . . . . . . 198

Page 19: Seismic Assessment of Reinforced Concrete Frame Structures

xxiv LIST OF FIGURES

Figure 5.3 General layout of the 4-storey RC building tested at ELSA (dimensions in metres) . . . . . . 204

Figure 5.4 Input accelerogram (Friuli-like) and elastic response spectra . . . . . . . . . . . . . . . . . . 212

Figure 5.5 Time histories of storey displacements, relative inter-storey drift, total storey-shear and respective peak value profiles for Low and High level tests . . . . . . . . . . . . . . . . 215

Figure 5.6 Shear-drift diagrams at each storey, for Low and High level tests. . . . . . . . . . . . . . . . 216

Figure 5.7 Time histories of dissipated energy and base shear - top displacement diagrams for Low and High level tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217

Figure 5.8 Detail of the 2nd storey experimental shear-drift diagram for the high level test . . . . . . . . 220

Figure 5.9 Spatial distribution of rotations in the critical zones, for the high level test. . . . . . . . . . . 221

Figure 5.10 Time histories of storey displacements, relative inter-storey drift, total storey-shear and respective peak value profiles for both configurations of infilled frame tests. . . . . . . . 223

Figure 5.11 Shear-drift diagrams at each storey for both configurations of infilled frame tests . . . . . . . 224

Figure 5.12 Time history of the imposed top displacement for the final cyclic tests . . . . . . . . . . . . . 227

Figure 5.13 Base shear - top displacement diagrams and curves of total deformation energy for the final cyclic tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229

Figure 5.14 Storey profiles of peak values of displacement, inter-storey drift and inter-storey shear for the final cyclic tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230

Chapter 6

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING . . . . . . . . . . . . . . . . . . . . . . . 243

Figure 6.1 Mesh for the structural analysis using flexibility global elements (dimensions in m) . . . . . . 245

Figure 6.2 Typical beam and column cross-sections for both external and internal frames. . . . . . . . . 249

Figure 6.3 Schematic reinforcement layout for the beams. . . . . . . . . . . . . . . . . . . . . . . . . . 250

Figure 6.4 Schematic reinforcement layout for the columns . . . . . . . . . . . . . . . . . . . . . . . . 251

Figure 6.5 Moment-curvature diagram for a column section. Comparison of trilinear curve and fibre analysis results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254

Figure 6.6 Diagonal strut model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261

Figure 6.7 Equivalence of element in anti-symmetric bending with cantilever elements . . . . . . . . . . 270

Figure 6.8 Yielding chord rotation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273

Figure 6.9 Estimation of ultimate chord rotation by curvature integration . . . . . . . . . . . . . . . . . 275

Figure 6.10 Example of inconsistency in energy splitting between element end sections: a) Beam and deformed shape, b) bending moments and c) curvature diagrams. . . . . . . . . 278

Figure 6.11 Storey displacement prescription and unloading to zero actuator forces . . . . . . . . . . . . 282

Page 20: Seismic Assessment of Reinforced Concrete Frame Structures

LIST OF FIGURES xxv

Figure 6.12 Introduction of the infill panel diagonal struts. . . . . . . . . . . . . . . . . . . . . . . . . . 283

Figure 6.13 Unloading of infilled frame configuration: a) removal of actuators and b), removal of infill panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284

Figure 6.14 Pushover analysis: inverted triangular force distribution. . . . . . . . . . . . . . . . . . . . . 285

Figure 6.15 Pushover analysis: base shear-top displacement and storey shear-drift diagrams. . . . . . . . 285

Figure 6.16 0.4S7 test. Static analysis versus experimental storey results . . . . . . . . . . . . . . . . . 288

Figure 6.17 0.4S7 test. Static analysis versus experimental results . . . . . . . . . . . . . . . . . . . . . 289

Figure 6.18 1.5S7 test. Static analysis versus experimental storey results . . . . . . . . . . . . . . . . . 292

Figure 6.19 1.5S7 test. Static analysis versus experimental results . . . . . . . . . . . . . . . . . . . . . 293

Figure 6.20 1.5S7 test - Static analysis. Spatial distributions of peak values . . . . . . . . . . . . . . . . 294

Figure 6.21 Uniformly infilled test. Static analysis versus experimental results . . . . . . . . . . . . . . . 298

Figure 6.22 Soft-storey test. Static analysis versus experimental results. . . . . . . . . . . . . . . . . . . 299

Figure 6.23 Shear-drift diagrams for the Ductility 3 phase of final test. Effects of considering or neglecting infilled frame tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302

Figure 6.24 Inter-storey drift profiles for 1.5S7, soft-storey and final Duct. 3 tests . . . . . . . . . . . . . 303

Figure 6.25 Final cyclic test, Ductilities 5 and 8: first and second storey shear-drift diagrams . . . . . . . 304

Figure 6.26 Final cyclic test: total energy diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305

Figure 6.27 Final cyclic test Duct. 8: results for modified unloading stiffness degradation . . . . . . . . . 306

Figure 6.28 Final cyclic test: profiles of peak values of storey shear . . . . . . . . . . . . . . . . . . . . 307

Figure 6.29 Influence of assumed displacements different from the applied ones . . . . . . . . . . . . . . 308

Figure 6.30 0.4S7 test. Dynamic analysis with 1.8% viscous damping versus experimental storey results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313

Figure 6.31 0.4S7 test. Dynamic analysis with 1.8% viscous damping vs. experimental results . . . . . . 314

Figure 6.32 0.4S7 test. Dynamic analysis with no viscous damping versus experimental results . . . . . . 315

Figure 6.33 1.5S7 test. Dynamic analysis with 1.8% viscous damping vs. experimental results . . . . . . 317

Figure 6.34 1.5S7 test. Dynamic analysis with 1.8% viscous damping and modified pinching versus experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318

Figure 6.35 1.5S7 test. Dynamic analysis with zero viscous damping versus experimental storey results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320

Figure 6.36 1.5S7 test. Dynamic analysis with no viscous damping versus experimental results . . . . . . 321

Figure 6.37 Total input energy for experimental and numerical analysis . . . . . . . . . . . . . . . . . . 323

Figure 6.38 Relative absorbed energy for experimental and numerical analysis. . . . . . . . . . . . . . . 324

Page 21: Seismic Assessment of Reinforced Concrete Frame Structures

xxvi LIST OF FIGURES

Figure 6.39 Member discretization for fixed length plastic hinge (F.H.) analysis . . . . . . . . . . . . . . 326

Figure 6.40 0.4S7 test. Static analysis with F.H. modelling . . . . . . . . . . . . . . . . . . . . . . . . . 329

Figure 6.41 1.5S7 test. Static analysis with F.H. modelling . . . . . . . . . . . . . . . . . . . . . . . . . 330

Figure 6.42 0.4S7 test. Dynamic analysis with F.H. modelling and 1.8% viscous damping . . . . . . . . . 332

Figure 6.43 0.4S7 test. Dynamic analysis with F.H. modelling and zero viscous damping . . . . . . . . . 333

Figure 6.44 1.5S7 test. Dynamic analysis with F.H. modelling and zero viscous damping . . . . . . . . . 334

Chapter 7

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 . . . 341

Figure 7.1 Basic configurations of the eight storey trial cases (PREC8) . . . . . . . . . . . . . . . . . . 347

Figure 7.2 Artificial accelerograms (S1...S4) and response spectra (5% damping) fitting the EC8 response spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350

Figure 7.3 Structural systems of planar frame associations . . . . . . . . . . . . . . . . . . . . . . . . . 352

Figure 7.4 Approximation of the post-yielding branch of the (M-ϕ) diagrams in columns . . . . . . . . . 355

Figure 7.5 Beam section moment-curvature diagrams of all cases with configuration 6 in direction YY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357

Figure 7.6 Column section moment-curvature diagrams of all cases with configuration 6 in direction YY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 358

Figure 7.7 Global overstrength. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360

Figure 7.8 Base shear - top displacement curve for the C2_15L case, direction XX. Definition of global yielding force. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361

Figure 7.9 Global hardening factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363

Figure 7.10 Cracking pattern: Configuration 6, Direction X under earthquake S1 for intensity 1.0 and 2.0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366

Figure 7.11 Positive rotation ductility pattern: Configuration 6, Direction X under earthquake S1 for intensity 1.0 and 2.0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367

Figure 7.12 Damage pattern: Configuration 6, Direction X under earthquake S1 for intensity 1.0 and 2.0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368

Figure 7.13 Damage pattern: Configuration 2, Direction X under earthquake S1 for intensity 1.0 and 2.0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369

Figure 7.14 Column rotation ductility profiles (maxima) . . . . . . . . . . . . . . . . . . . . . . . . . . . 372

Figure 7.15 Beam rotation ductility profiles (maxima) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373

Figure 7.16 Column damage profiles (maxima) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374

Figure 7.17 Beam damage profiles (maxima). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375

Page 22: Seismic Assessment of Reinforced Concrete Frame Structures

LIST OF FIGURES xxvii

Figure 7.18 Total drift (%) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377

Figure 7.19 Maximum inter-storey drift (%) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378

Figure 7.20 Sensitivity coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380

Figure 7.21 Maximum damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381

Figure 7.22 Global (average) damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382

Figure 7.23 Local probability of failure at the hinge level . . . . . . . . . . . . . . . . . . . . . . . . . . 384

Figure 7.24 High and medium seismicity hazard curves . . . . . . . . . . . . . . . . . . . . . . . . . . . 387

Figure 7.25 Vulnerability curve fitting to non-linear analysis results . . . . . . . . . . . . . . . . . . . . 388

Figure 7.26 Bounds of annual probability of failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391

Appendix A

Linear Elastic Timoshenko Beam Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421

Figure A.1 Distribution of flexibility properties along the element . . . . . . . . . . . . . . . . . . . . . 423

Appendix B

Trilinear Model Details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431

Figure B.1 Hysteretic rules of the trilinear model. General loading path . . . . . . . . . . . . . . . . . . 431

Figure B.2 Trilinear model. The pinching effect and interior cycles . . . . . . . . . . . . . . . . . . . . 433

Figure B.3 Trilinear model. Strength deterioration rule . . . . . . . . . . . . . . . . . . . . . . . . . . . 434

Appendix C

Internal Force Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437

Figure C.1 Element applied loads in the non-linear bending plan; simplified notation . . . . . . . . . . . 440

Appendix D

The Event-to-Event Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445

Figure D.1 The event-to-event scheme for stiffness based problems . . . . . . . . . . . . . . . . . . . . 445

Page 23: Seismic Assessment of Reinforced Concrete Frame Structures
Page 24: Seismic Assessment of Reinforced Concrete Frame Structures

LIST OF TABLES

Chapter 3FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL. . . . . . 63

Table 3.1 Definition of cracking section abscissas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Chapter 4NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION . . . . . . . . . . 141

Table 4.1 Mechanical properties of steel and concrete of LNEC and KT beams . . . . . . . . . . . . . 178

Chapter 5THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA . . . . . . . . . . . . . . . . . . . . 191

Table 5.1 Mean cube (fcm,cub) and cylinder (fcm) compressive and tensile (fctm) strengths of concrete . . 205

Table 5.2 Mean tensile properties of steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206

Table 5.3 Frequencies (Hz) for all testing cases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214

Chapter 6ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING. . . . . . . . . . . . . . . . . . . . . . . 243

Table 6.1 Floor mass values and vertical loads on beams . . . . . . . . . . . . . . . . . . . . . . . . . 248

Table 6.2 Mean tensile properties of steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253

Table 6.3 Numerical simulations performed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280

Table 6.4 Structural frequencies. Comparison of measured values with those calculated by flexibility discretization310

Chapter 7SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8. . . 341

Table 7.1 Trial cases, design behaviour factors and earthquake intensities . . . . . . . . . . . . . . . . 348

Table 7.2 Member cross-sectional dimensions (m) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349

Table 7.3 Floor masses. Adopted values and design values (in brackets) . . . . . . . . . . . . . . . . . 353

Table 7.4 Frequencies (Hz) for all cases (design values in brackets) . . . . . . . . . . . . . . . . . . . 353

Table 7.5 Design base shear force ratio to structure weight (seismic coefficient) . . . . . . . . . . . . . 359

Table 7.6 Overstrength factors at yielding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362

Page 25: Seismic Assessment of Reinforced Concrete Frame Structures

Chapter 1

INTRODUCTION

1.1 General

Earthquake Engineering is an ever challenging research field, due to the wide variety of issues

involved. These can be as different as the seismic event generation and the characterization of

seismic motion at a given site, or the dynamic behaviour of structures and their non-linear

response, or the definition of measures to assess the structural reliability under seismic events.

In this context, a given structure can be regarded as a sort of “operator” mapping the seismic

input action into effects that are compared against some performance limits which are estab-

lished according to socially and economically acceptable costs.

It follows that, regardless of the seismic input and the response reliability analysis, an adequate

knowledge of the structural dynamic behaviour appears of major importance for the seismic

assessment of structures. The peculiar features of the reinforced concrete (RC) behaviour, par-

ticularly under cyclic loading, introduce increased levels of complexity in the non-linear struc-

tural response, for which significant experimental testing would be desirable, with particular

interest on large or full scale specimens.

Because of the extremely high cost of large scale tests, either dynamic or pseudo-dynamic,

they are often restricted to a few cases that may be considered representative of certain types of

structures. Thus, recourse has to be made to numerical modelling of the structural hysteretic

response by means of non-linear behaviour models, which are calibrated on the basis of widely

available results from experiments on single reinforced concrete members or sub-assemblages;

such calibration can be complemented, or even checked, by comparing the numerical results of

the structural response with available experimental evidence from large or full scale tests.

Page 26: Seismic Assessment of Reinforced Concrete Frame Structures

2 Chapter 1

The importance of modelling the non-linear hysteretic behaviour of RC structures arises from

the present seismic design philosophy, according to which an adequate seismic behaviour of a

given structure should fulfil the following requirements:

• under the action of low intensity earthquakes (A), that may occur several times during the

structure lifetime, no structural or non-structural damage should be detected;

• the response to moderate intensity seismic events (B), typically those occurring once in the

structure lifetime, may lead to slight (but insignificant) structural damage associated with

visible non-structural damage;

• for major earthquakes (C), thus with very low probability of occurrence during the structure

lifetime, significant but repairable structural damage can be expected as long as no partial

nor total collapse of the structure occurs.

From the above, it follows that the structural response will be distinct for each level of seismic

intensity. Typically, for the intensity level A an essentially linear elastic behaviour shall be

expected, while for intensity B the structure will exhibit apparent cracking on structural mem-

bers and, possibly, some incipient and localized yielding. On the other hand, for earthquakes of

intensity C, significant inelastic behaviour is likely to develop in a generalized fashion

throughout the critical zones of the structure where hysteretic dissipation capability must be

provided to cope with the seismic induced energy.

According to that seismic design philosophy, the intensity levels A and B may be assigned lin-

ear elastic behaviour, preferably associated with dynamic analysis for the seismic response

prediction. However, recent studies point out the need to account for the non-linear effects aris-

ing from cracking of reinforced concrete members, even for such levels of seismic action, in

order to obtain adequate estimates of the structural response (Fardis and Panagiotakos (1997),

Calvi and Pinto (1996)); indeed, this corresponds to the usual problem of deciding which stiff-

ness shall be used for linear elastic analysis (uncracked, cracked or some intermediate state).

The analysis of seismic effects arising from earthquakes of intensity level C should be ideally

performed by means of non-linear behaviour models, able to account for the exploitation of

high ductility (generically defined as the inelastic deformation capacity without significant loss

of strength) and for the capacity of energy dissipation through stable hysteretic mechanisms.

However, such type of analyses (particularly in the dynamic context) are not yet possible in

Page 27: Seismic Assessment of Reinforced Concrete Frame Structures

INTRODUCTION 3

current seismic design practice because of their complexity. Thus, seismic design codes allow

linear elastic analysis to be performed for a seismic action q times lower than the design one,

where q is the so-called behaviour factor, through which the non-linear effects are approxi-

mately taken into account. For RC structures the q-factor typically varies between 2 and 5, and

basically it means that the structure must provide the strength for forces q times lower than

those resulting from the linear elastic analysis with the design earthquake and, simultaneously,

have the capacity of accommodating inelastic deformations of the order of those obtained in

that elastic analysis. Such inelastic deformation capacity shall be guaranteed in the critical

zones bound to behave non-linearly, which have to be provided with a certain amount of avail-

able ductility; the larger is the q-factor the more available ductility is required in those critical

zones.

Nevertheless, such simplified analysis require some issues to be investigated, viz the calibra-

tion of q factors and their relation with ductility required in dissipative zones of the structure,

as well as the influence of the structure geometry and the distributions of mass and stiffness on

the response to high intensity earthquakes. As suggested by Fardis (1991), q-factor calibration

can be made through parametric studies aiming at checking whether or not a structure will be

able to accommodate permanent, yet repairable, damage under a seismic input q times stronger

than the one used to obtain its design load-effects by linear elastic analysis.

Parametric studies in the seismic analysis context typically require large computational effort

as a result of time domain analyses, generally for several earthquakes characterizing the seis-

mic input and, quite often, for several intensities of the seismic action. Thus, adequate analysis

models have to be carefully chosen, through which the various non-linear behaviour mecha-

nisms can be simulated with acceptable realism, at a manageable computational cost.

For RC frame structures, constituting the structural type the present work refers to, member

models are deemed particularly suitable to carry out seismic non-linear analysis, since they

provide the best compromise of response detail, efficiency and simplicity. Each member

(beam, column) is typically discretized by a single element in which the behaviour is most

often controlled by phenomenological models at certain critical sections; a description of dam-

age and non-linear effects along the member can be obtained, whose detail depends on the for-

mulation underlying the element model. Therefore, besides the global structure behaviour,

member models provide some insight into the member response and major goals of parametric

Page 28: Seismic Assessment of Reinforced Concrete Frame Structures

4 Chapter 1

studies can be achieved, such as the local ductility or damage assessment and their spatial dis-

tribution throughout the structure.

Many member models have been proposed to date, some of them assuming the inelastic behav-

iour lumped in “point hinges” at member ends, while others consider the inelasticity spreading

along the member in order to better describe the actual behaviour. Most member models were

cast in the framework of the classical stiffness method using cubic hermitian polynomials to

approximate the element displacement field as for linear elastic elements. However, for distrib-

uted inelasticity elements this approach deviates from the actual behaviour because the cubic

representation of displacements becomes less adequate as the member stiffness distribution

changes due to inelasticity development. Upon recognition of this problem, some attempts

were proposed for improved representation of internal deformations along the element but the

most consistent approaches were developed in the flexibility formulation context (Ciampi and

Carlesimo (1986), Taucer et al. (1991), Spacone et al. (1992)).

In flexibility based elements use is made of force interpolation functions strictly derived from

equilibrium conditions. This constitutes a major advantage over stiffness based elements

because such functions are exact regardless of the damaged state of the member and, addition-

ally, they can be straightforwardly derived for frame elements. However, there is a price to

pay: since no displacement interpolation functions are available, the element state determina-

tion cannot be directly performed. An internal iterative scheme is thus required to compute the

element resisting forces associated with the imposed displacements at element nodes; during

that iterative scheme the state determination is performed at a few control sections and both the

flexibility matrix and the displacements of the element are obtained from integration of the

section flexibility and deformation distributions, respectively. The scheme is driven by gradual

elimination of residual displacements while strictly preserving equilibrium along the element.

The formulation is general in the sense that, provided the force interpolation functions are

defined, it can be used with any type of section modelling technique, viz the fibre model and

global section models of phenomenological or differential type; in addition, it is very suitable

to accommodate associations in series of several elements behaving non-linearly, each

accounting for a specific source of non-linearity.

In a few words, not only the formulation of flexibility based elements appears a rather promis-

Page 29: Seismic Assessment of Reinforced Concrete Frame Structures

INTRODUCTION 5

ing and elegant technique for frame analysis, as it also brought back to light the somewhat

“forgotten” role of the force method in structural analysis.

1.2 Objectives

With the advent of flexibility based elements a few elements were proposed with considerable

level of sophistication and quite good adequacy for describing the non-linear hysteretic behav-

iour of frame members. Particular reference is made to the fibre beam-column element (Taucer

et al. (1991)), the first consistently developed and implemented in the flexibility formulation

framework; however, despite the model ability to describe complex hysteretic behaviour, its

use may easily become prohibitive in seismic analysis due to the extremely high computational

and data management cost.

Other flexibility based models were proposed by Spacone et al. (1992) and Filippou et al.

(1992), in which global section constitutive laws are adopted to control the RC non-linear

behaviour. In the former, a differential constitutive relationship is used to monitor several pre-

defined control sections; in the latter, a phenomenological behaviour law is adopted and the

spread of post-yielding inelasticity is considered through progressive development of plastic

zones. However, neither of them accounts for the cracking spread along the member because

the non-linear response is assumed to start from the cracked stage.

In the present work, special attention was devoted to the development and implementation of a

new flexibility based element model, which is robust, efficient and economic, both in discreti-

zation and computation time requirements, while being able to adequately describe the cyclic

response of RC members for different behaviour stages - uncracked, cracked and yielded.

The proposed element is particularly suitable for the analysis of members under cyclic defor-

mation reversals in bending, possibly combined with low levels of axial load. Bi-axial bending

is considered, though assuming non-linear behaviour only for one bending plane; the other is

assumed to behave linearly. Moreover, the actual behaviour is approximated by uncoupling the

effects of axial force from the bending behaviour, as a result of the assumption of low axial

force values.

In line with this new element proposal, the following objectives were envisaged:

• To develop the element model in the framework of the flexibility formulation, approxi-

Page 30: Seismic Assessment of Reinforced Concrete Frame Structures

6 Chapter 1

mately following the basic steps for the element state determination as proposed in previous

pioneering works (Taucer et al. (1991), Spacone et al. (1992)). The element model should

fulfil a basic discretization feature of member models, i.e. one structural member is to be

modelled by only one element, thus requiring due consideration of element applied loads

(distributed or concentrated). The non-linear behaviour should be controlled at the global

section level with a multi-linear phenomenological model and the number of control sec-

tions should be the minimum possible at each load step.

• To implement the element model in a general purpose computer code for the non-linear

static and dynamic analysis of structures.

• To develop and to implement in the same code, a new algorithm for the definition of the

basic curves of the section model for the most common cases of rectangular and T-shape RC

sections.

• To apply the proposed element to the numerical analysis of a four-storey full-scale RC

building pseudo-dynamically tested under two different earthquake loading levels and

quasi-statically loaded up to failure by means of cyclic tests of increasing intensity. This

should serve to check the model ability to simulate the global structure behaviour through-

out distinct stages, both in quasi-static or dynamic conditions. On the other hand, an addi-

tional insight on the structure response could be obtained, in particular concerning the

distribution and quantification of damage in the various members.

Another major aim of the present work is to provide some contribution to the seismic behav-

iour assessment of reinforced concrete frame structures. In view of the forthcoming approval

of Eurocode 8 (EC8) as a european standard, particular concern is devoted to assess the seis-

mic performance of structures designed according to that code. Therefore, as part of a euro-

pean-wide project of “Prenormative Research in Support of Eurocode 8” (PREC8) set up by

the European Commission and National Authorities, the numerical seismic analysis of some

EC8 designed building structures (9 out of a total of 26) was sought in order to assess “the

interrelation between a number of design parameters used in EC8, which, in a combined form,

influence the non-linear behaviour of structures subjected to earthquake motion” (Carvalho et

al. (1996), Pinto and Calvi (1996)).

1.3 Chapter organization

Page 31: Seismic Assessment of Reinforced Concrete Frame Structures

INTRODUCTION 7

The present Thesis is organized in eight Chapters and five Appendices, covering the descrip-

tion of the new flexibility based element development and implementation (Chapters 2 to 4)

and its application to the seismic analysis of RC frame structures (Chapters 5 to 7). The

Appendices include further details on specific topics which are deemed unnecessary for an

adequate reading of the global development and implementation schemes.

Following the introduction, Chapter 2 focuses on the modelling techniques of RC frames

under earthquake horizontal actions. The typical features of structural member response to

monotonic and cyclic loading are recalled in order to highlight the basics underlying the deri-

vation of phenomenological models. Numerical tools for non-linear modelling of reinforced

concrete buildings are briefly reviewed and special attention is given to member models; the

evolution of distributed inelasticity member models and the most relevant phenomenological

hysteretic models developed to date are addressed. The formulation of the most recent flexibil-

ity based elements is recognized as rather promising and suitable for frame member modelling,

for which it has been chosen as the framework for the element model development to be pur-

sued herein. Therefore, the general flexibility formulation for frame elements is addressed in

contrast with the classical stiffness method and the element state determination is identified as

the most critical task because displacement shape functions are not available. The iterative

scheme required for the element state determination is then presented as proposed in previous

pioneering works; no specific requirement is considered for the section model and due account

is taken of distributed or concentrated loads that may be applied to the element.

Chapter 3 is devoted to the full description of the new flexibility element development. A glo-

bal section constitutive law is adopted, consisting of a multi-linear step wise model based on

trilinear envelope curves and a number of hysteretic rules; it is basically a Takeda-type model

controlling the moment-curvature relationship, which accounts for the cracking and yielding

stages and for other typical phenomena of reinforced concrete, viz the stiffness and strength

degradation as well as “pinching”.

Two types of control sections are considered: the fixed control sections, consisting of the two

element end and one mid-span sections, and the moving sections, accounting for the so-called

yielding, cracking and null-moment sections. For a given load step, yielding sections define

the transition between sections having already yielded and those still in the pre-yielding range;

cracking sections establish the transition between cracked and uncracked sections, and null-

Page 32: Seismic Assessment of Reinforced Concrete Frame Structures

8 Chapter 1

moment sections are considered to allow for possible different behaviour in positive and nega-

tive bending directions. The set of fixed and moving sections defines distinct zones (yielded,

cracked and uncracked) which change during the loading history and allow to perform the inte-

grations inherent in the element state determination. Although an existing Takeda-type model

is used, a modification has to be considered consisting in a special transition from uncracked to

full-cracked behaviour in order to provide an approximate control of cracked zones.

The inclusion of uniformly distributed or concentrated element applied forces is discussed and

the development is cast for that purpose. However, the consideration of distributed forces was

found to require a rather cumbersome implementation scheme for the present stage of the algo-

rithm development; thus, it was decided to approximately account for the distributed forces by

means of an equivalent concentrated force applied in the mid-span section.

The element state determination is carried out by means of an internal element iterative

scheme, during which the flexibility distributions are defined in the distinct zones in order to

obtain the element flexibility matrix; the moving nature of cracking sections, where a curva-

ture discontinuity is considered, has been found to introduce additional flexibility contribu-

tions which are duly taken into account in the element matrix. The restricted number of fixed

control sections and the use of moving ones, requires particular attention to the evaluation of

displacement residuals which are computed through a different scheme than the one used in the

general flexibility formulation. Furthermore, an event-to-event scheme inside the element iter-

ative process has been found adequate to carefully control the non-linear behaviour of the

yielded end-zones. In addition, some convergence problems related with the cracking transi-

tion required the inclusion of a line-search scheme during internal iterations.

The basic steps of the general non-linear algorithm for static and dynamic analysis are summa-

rized, outlining the full sequence of steps of the element state determination. At the structure

level, the Newton-Raphson method is used to solve the non-linear equation system and the

classic Newmark scheme is adopted for the time integration of the dynamic equilibrium equa-

tions.

Chapter 4 describes the main interventions for the implementation of the new element model

in the general purpose computer code CASTEM2000. Because this object-oriented code

exhibits particular features, very different from traditional specific purpose programs, a brief

Page 33: Seismic Assessment of Reinforced Concrete Frame Structures

INTRODUCTION 9

review of CASTEM2000 basics is given, illustrated by a rather simple structural analysis prob-

lem; the major concern is to introduce some key concepts such as objects, commands, opera-

tors and procedures that integrate the code environment.

As an auxiliary tool to perform a pre-processing task, a new algorithm was developed and

implemented to define the trilinear envelope curves for rectangular and T-shape sections.

Instead of making a fibre-type analysis, which can be regarded as a general technique for bend-

ing analysis of RC sections, an algorithm was specifically designed for rectangular and T-

shape sections based on a systematization of the internal equilibrium conditions for such type

of sections. Realistic material models are used, viz a bilinear one for steel (with strain harden-

ing) and a parabolic/linear-softening/residual-plateau model for concrete; confined and uncon-

fined zones are duly distinguished in the section. The cracking point in the moment curvature

envelope is defined by classic closed form expressions, while the remaining turning points

(yielding and ultimate) are obtained by a unified process, though with different criteria for the

point definition.

A few validation tests at the single element level are also reported in Chapter 4, by comparison

with experimental results from RC members tested under monotonic and cyclic loading; the

main features of the model response can be identified and possible limitations can be detected

which should be taken into account when analysing complete structures.

Chapter 5 presents the experimental tests in the four-storey full-scale RC building carried out

in the European Laboratory for Structural Assessment (ELSA) of the Joint Research Centre

(JRC), Ispra, Italy, where the author has developed the present work. Included in a wider

research programme involving several european partners, this testing campaign aimed at a

comparison between the actual and the expected design behaviour of a high ductility structure

designed according to Eurocodes 2 and 8. The experimental activity consisted of preliminary

tests (free vibration and stiffness) for dynamic characterization of the structure, followed by

unidirectional seismic tests pseudo-dynamically performed in the bare structure for two inten-

sity levels (0.4 and 1.5) of the reference earthquake, artificially generated to fit the EC8 elastic

spectrum used in the design. Further seismic tests at 1.5 intensity were carried-out in the struc-

ture with unreinforced masonry infills in the external frames, both in a totally infilled configu-

ration and in a partially infilled one where the first storey was kept bare in order to simulate

soft-storey conditions. Finally, a quasi-static cyclic test was performed in the bare structure up

Page 34: Seismic Assessment of Reinforced Concrete Frame Structures

10 Chapter 1

to a near-failure stage.

A brief description of the pseudo-dynamic testing technique is included and a few comments

are made on the structure design and layout, the characterization tests of material properties

and the reduced scale member tests prior to the full-scale ones. The results of tests on the com-

plete structure are thoroughly described in order to provide a global view of the structural per-

formance throughout different behaviour stages; this was deemed important to support the

discussion of numerical simulations in the subsequent chapter.

Chapter 6 presents and discusses the application of the proposed element to the numerical sim-

ulation of the full-scale tests performed in the four-storey structure at the ELSA laboratory.

Because these tests covered different behaviour stages (pre-yielding and post-yielding with

increasing ductility levels up to failure) and due to their essentially static nature (from a strict

experimental standpoint), they provide an excellent means of calibration and assessment of

numerical models, through either static simulations with the experimentally imposed displace-

ment or dynamic calculations with the accelerogram used in the experiment. All the performed

tests (except the preliminary ones) are simulated, both in the bare and the infilled structure

configurations. Therefore, the presentation and discussion of modelling assumptions includes

also references to the infill panel modelling which is accomplished by pairs of diagonal struts

ruled by uniaxial force-deformation models whose parameters are identified by means of bi-

dimensional refined analysis.

For its importance in the seismic assessment of structures, the damage quantification is briefly

addressed, referring and discussing some available proposals of damage indices. The widely

used Park and Ang index is adopted to calculate damage via the chord rotation at each element

end section. Particular attention is devoted to the quantification of parameters required in that

index, viz the ultimate rotation and the dissipated energy by hysteresis.

The numerical simulations are divided into static and dynamic ones. The former are performed

by applying the displacements actually imposed in the experiment and provide structural

strength responses to be compared with experimental results without involving additional

dynamic effects. Static simulations are carried out for all tests (the seismic ones in both bare

and infilled structures, and the final cyclic test) and particular care is taken to follow the actual

test sequence as close as possible; the real conditions of load application, including the unload-

Page 35: Seismic Assessment of Reinforced Concrete Frame Structures

INTRODUCTION 11

ing phases between tests, are simulated by adequate adaptation of boundary conditions, for

which the modularity and object-oriented features of CASTEM2000 proved to be rather suita-

ble. Static pushover analyses are performed under inverted triangular distribution of monoton-

ically increasing forces in order to provide estimates for the maximum base-shear and the

global yielding displacement. The quality of static simulation results is discussed through com-

parison with the experimental ones, mainly in terms of storey shear forces, shear-drift dia-

grams and dissipated energy, as well as rotations in beams. Additional analytical response

variables, such as rotation ductility and damage, are also presented even though no experimen-

tal counterpart is available. Results are sequentially analysed, first for the bare structure seis-

mic tests, then for the infilled structure and finally for the quasi-static cyclic test. Dynamic

simulations are restricted to the bare structure seismic tests and preceded by the comparison of

calculated and experimentally measured structural frequencies for several testing stages.

Results of dynamic calculations are compared with experimental ones, mainly in terms of sto-

rey displacements, shear forces and shear-drift diagrams; evolutions of the total dissipated

energy are also compared. The viscous damping characterization is discussed at length for both

seismic tests (0.4 and 1.5) in view of the dissipation capacity of the model for the different

behaviour stages and some comments are made on the validity of comparing analytical and

experimental dissipated energy in the dynamic context.

Finally the proposed flexibility element modelling strategy is compared with a more classical

one, in which each member is discretized by one linear elastic element and two non-linear ele-

ments to simulate the plastic hinge zones at the member ends. Besides the presentation of the

assumptions for this modelling strategy, results are compared for the seismic tests on the bare

structures, through both static and dynamic simulations; mainly the quality of results is com-

pared (by means of the same response variables as before), while no specific comparison is

made concerning model efficiency due to the different discretization needs in the two model-

ling options.

Chapter 7 is devoted to the presentation and discussion of the numerical seismic assessment of

some RC frame structures included in the PREC8 programme, resulting from our activities in

the JRC. Because the major scope is concerned with the implications of EC8 provisions on the

seismic behaviour of building structures, a brief review of EC8 basics is included mainly to

recall the principles of design philosophy that must be kept in mind for result analysis. The

analysed structures have two basic configurations, one regular and another irregular, both with

Page 36: Seismic Assessment of Reinforced Concrete Frame Structures

12 Chapter 1

eight storeys and a rectangular plan of 15x20 m2. Each configuration was designed for two

design accelerations (0.15g and 0.30g) and different ductility classes; in addition, the irregular

configuration was also designed using the simplified static analysis method allowed in EC8.

Nine trial case structures are modelled with the proposed flexibility element, following model-

ling assumptions similar to those of the building analysed in Chapter 6.

Non-linear dynamic planar analyses are performed for each trial case in both horizontal direc-

tions, under the action of four accelerograms (artificially generated to fit the EC8 elastic spec-

trum) with a number of increasing intensities. Additionally, pushover static analyses are

carried out to allow the quantification of structural global overstrength in terms of the base-

shear force, whose results are extensively discussed and compared between the trial cases. The

spread of seismic effects on the structures is analysed by means of cracking patterns and spatial

distributions of ductility and damage. The overall structural response is described by global

parameters such as the total drift, inter-storey drift and damage (maximum and global aver-

age). Where pertinent, comparisons are made with EC8 limits (namely for the serviceability

limit state), though the most systematic comparative analysis is made between trial cases

according to the different design accelerations and ductility classes.

It is also included an attempt of system reliability analysis of the various structures, aiming at

estimates of probability of failure, mainly for comparative purposes between the different trial

cases rather than an absolute evaluation of the structural safety. The quantification of probabil-

ities of failure is based on the damage values in the critical zones (plastic hinges); the adopted

methodology and assumptions for this complex topic are described on the basis of previous

studies addressing the computation of local hinge probabilities of failure, the probabilistic

quantification of the seismic intensity, the damage capacity characterization and the computa-

tion of estimates for the system probability of failure. Upper and lower bounds of the latter are

calculated and compared between the analysed structures aiming at an assessment of design

parameter (e.g. ductility class, design acceleration) influence on the structural safety.

Finally, the most relevant results and conclusions of this work are summarized in Chapter 8,

where suggestions for future research are also pointed out in line with the developments made

in this thesis.

Page 37: Seismic Assessment of Reinforced Concrete Frame Structures

Chapter 2

MODELLING OF REINFORCED CONCRETE FRAME

STRUCTURES - STATE OF THE ART

2.1 Introduction

It is currently well established the need of adequate non-linear models to account for the hys-

teretic behaviour of RC frame structures. The major purposes are related with the calibration of

some design parameters, such as q-factors and the inherent ductility requirements in structural

dissipative zones, as well as possible irregularities related with both the structure geometry and

the distributions of mass and stiffness. These calibration studies basically aim at giving support

to simplified procedures (e.g. linear elastic analysis using q-factors, rules for irregularity clas-

sification, detailing provisions to assure a certain ductile capacity), constituting sets of practi-

cal tools for current seismic design.

Under high intensity earthquakes, the structural behaviour of buildings is usually controlled by

its resistance to horizontal actions, from which the inelastic behaviour concentrates in the end

zones of structural members. However, besides other reasons related with design options and

structural configurations, the spatial distribution of such zones is also affected by the possible

preponderance of gravity loads, which, therefore, shall be taken into account.

Moreover, recent studies which point out the damaging potential of vertical ground motion

(Elnashai and Papazoglou (1995), Papazoglou and Elnashai (1996)), have shown in some cases

the lack of conservatism of vertical earthquake forces as estimated by current code spectra. In

addition, proposals for the inclusion of vertical motion effects in seismic design can be found

in the literature (e.g. Elnashai and Papazoglou (1997)). Despite this recent evidence and the

problem relevance for the seismic response of structures, particularly in near-source regions,

Page 38: Seismic Assessment of Reinforced Concrete Frame Structures

14 Chapter 2

the present work will focus only on the effects of the horizontal seismic components.

The scope of the present chapter is to provide a brief, yet comprehensive, description of mod-

elling techniques for the non-linear hysteretic behaviour of reinforced concrete frame struc-

tures, cast in the form of a historical review, from which the present state-of-the-art can be

addressed. Thus, the main features of structural member behaviour under monotonic and cyclic

loading are first recalled in 2.2, as they constitute the basic evidence for derivation of most

phenomenological behaviour models.

Numerical modelling strategies for reinforced concrete buildings are discussed in 2.3, with

particular emphasis on member models, which can provide sufficient detail of structural

response while keeping with manageable algorithms of analysis. Some of the most relevant

phenomenological hysteretic models are referred and the evolution of distributed inelasticity

member models is described, since, by contrast with point-hinge models, they allow to closely

follow the actual force distributions and the stiffness modifications along the member.

Among the most recently developed member models, the flexibility based ones appear as a

very elegant and suitable option for frame member modelling. Indeed, the flexibility formula-

tion allows to account for inelasticity spread in a quite natural way, while keeping complete

freedom for the choice of section behaviour model. In view of the member model developed in

the present work, which belongs to the flexibility based model family, the general flexibility

formulation for beam-column elements is described with some detail in 2.4, yet mainly based

on previous works available in the literature.

2.2 Hysteretic behaviour of reinforced concrete members

The global behaviour of a given structure reflects the behaviour of its members and corre-

sponding interconnections. Hence, experimental evidence from tests on single reinforced con-

crete members is of major importance for understanding the complete structure behaviour,

under either monotonic or cyclic loading conditions. Even if experimental tests do not exactly

match the actual conditions of members in a real structure, the source of relevant non-linear

phenomena remains essentially the same and, therefore, experimental results provide

extremely valuable information on non-linear response mechanisms.

In this context, the present section deals with the presentation of the main features of rein-

Page 39: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 15

forced concrete member behaviour, as they are commonly found in experimental tests of single

elements (e.g. cantilever or simply supported beams or columns) or beam-column sub-assem-

blages. Although widely known within the seismic (or generally cyclic) structural behaviour

context, such features are briefly recalled herein for completeness purposes in view of the sub-

sequent sections. Excessive detail is avoided, in particular that related with the specific behav-

iour of constitutive materials (steel and concrete) and their interaction, since they can be found

in standard literature (Park and Paulay (1975), Paulay and Priestley (1992)) or in available

broad reviews (Coelho (1992), Costa (1989)). Furthermore, only uniaxial bending response is

considered as it exhibits the most relevant behaviour features.

For adequate understanding of the hysteretic behaviour, the member response under monotonic

loading shall be addressed first. Two typical behaviour types are usually found: that of beam

members, hence with null or rather low level of axial force, and that of columns, normally with

non-negligible axial force level.

Figure 2.1 shows illustrative global response diagrams (e.g. tip force-displacement of a canti-

lever specimen) for both behaviour types, where, after an initial quasi-linear branch up to the

cracking force , another branch follows in correspondence with the cracked behaviour of

a certain zone along the member. In the vicinity of the highly non-linear response range (i.e.,

nearby the yielding threshold ), the behaviour is very different in beams and columns: the

former case is mainly controlled by the steel behaviour (for which a schematic steel stress-

strain diagram is included), while the latter reflects the typical compressive behaviour of con-

crete due to the predominance of section compression forces required for axial force equilib-

rium. By increasing the axial force, the ultimate displacement reduces due to the decreasing of

the section ductile capacity; in turn, the yielding threshold may increase or reduce, depending

on whether the axial force is less or greater than the so-called balanced value which defines the

transition from yielding controlled by the tensile steel towards yielding dominated by the com-

pressed concrete.

Monotonic response diagrams are also found to constitute envelopes of cyclic response, which

makes them rather useful for the definition of analysis models. Examples of cyclic behaviour

are included in Figure 2.2, referring to quasi-static cyclic tests (Carvalho (1993)) on a uniform

section cantilever beam of 1.5m span (as illustrated) under applied tip-displacement sequences

and with null axial force.

Fc( )

Fy( )

Page 40: Seismic Assessment of Reinforced Concrete Frame Structures

16 Chapter 2

Figure 2.1 Typical global response diagrams of beams and columns for monotonic loading

These diagrams exhibit the most typical features of the hysteretic response of reinforced con-

crete members; particularly referring to Figure 2.2-a), the following is highlighted:

• The load path 0-1 follows the monotonic response, clearly showing stiffness reductions due

to concrete cracking at C- and to reinforcement yielding at Y-.

• Unloading along 1-2 exhibits stiffness close to the initial cracked one (i.e., the slope of a

line connecting the origin with the yielding point Y-), indeed a behaviour similar to that of

the steel.

• The first reloading (inversion of force sign) follows the path 2-3, initially with a high stiff-

ness branch, after which the previously compressed but still uncracked section zones,

become progressively cracked, thus leading to stiffness reduction at point C+; this diagram

part develops as if the positive envelope were shifted backwards to the residual displace-

ment at point 2, therefore reaching yielding of bottom reinforcement still for negative dis-

placement at point Y+.

• Instead, for the subsequent reloading phases (6-7 or 10-11 in the positive direction and 4-5 or

8-9 in the negative one) apparent stiffness drops occur, which are generated by two distinct

(but mixed) phenomena:

- full-depth open cracks due to plastic elongation of steel, cause the section resisting

moments to be provided by a steel couple without any concrete contribution, which,

combined with bond deterioration between concrete and steel in the vicinity of crack

F

d

Fy

Fc fs

εs

Steel

F

d

Fy

Fc fc

εc Concrete

a) Beam (null or low axial force) b) Column

Page 41: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 17

lips, induce increased deformations to reach the same force of previous cycles;

- the typical Bauschinger effect exhibited by steel under cyclic loading, due to which, and

after post-yielding load reversal, a much pronounced non-linear response is obtained for

steel stress levels significantly lower than the monotonic yielding stress.

Figure 2.2 Cyclic global response examples (Carvalho (1993))

-50

-40

-30

-20

-10

0

10

20

30

40

50

-140 -120 -100 -80 -60 -40 -20 0 20 40 60 80 100 120 140

Displ (mm)

Forc

e (k

N)

b) Test S1-V5

-50

-40

-30

-20

-10

0

10

20

30

40

50

-140 -120 -100 -80 -60 -40 -20 0 20 40 60 80 100 120 140

Displ (mm)

Forc

e (k

N)

0 42

3

1

7

11

59

610

8

12

13

15

C-

C+

a) Test S1-V4

Y+

Y-

F+

0.2

4φ123φ12

0.3

1.5

δ+

Concrete: C25/30Steel: B500s

(Tempcore)

Page 42: Seismic Assessment of Reinforced Concrete Frame Structures

18 Chapter 2

• Reloading in the negative direction shows the so-called pinching effect: open cracks may

eventually close in the first reloading phase, hence activating again compressive stresses in

the concrete and increasing the stiffness as shown in the second reloading phase; crack clo-

sure is actually achieved in the negative direction because the larger top steel area can gen-

erate sufficient tensile forces to balance forces in the steel yielding in compression (which,

in turn, are necessary to compensate the previous plastic elongation). However, the same

does not occur for positive reloading and, consequently, cracks may never close again for

that direction. Furthermore, it is worth mentioning that, in symmetrically reinforced sec-

tions, due to the plastic elongation of steel, the load is likely to be carried very largely just

by the couple of steel forces; therefore, should significant axial compressive force be

present, higher compressive stresses are demanded and cracks tend to close earlier, which

means that the pinching effect may also become apparent in columns, particularly if large

excursions in the post-yielding range are likely to occur.

• For the same level of peak displacements, the resisting force is progressively decreasing as

a result of buckling of reinforcing bars, which started developing in cycle peaks 5 and 9 and

finally led to failure. This phenomenon of strength decay may also be triggered-off, or

aggravated, by less effective force transfer from steel to concrete due to bond deterioration

or by degradation of the concrete compressive strength, both causes arising from the cyclic

effect of crack opening and closing and of bars pulling in and out from concrete.

Another noteworthy feature is related with the unloading stiffness degradation, quite apparent

in Figure 2.2-b). The auxiliary dashed lines connecting the starting and end points of unloading

branches show decreasing slope for increasing inelastic excursions, in agreement with the pro-

gressive reduction of average stiffness of cycles for larger amplitude; actually, Figure 2.2-b)

shows reloading branches typically pointing at the maximum deformation reached in the previ-

ous cycle, which lead to that average stiffness reduction.

Finally, it must be pointed out that the above referred pinching effect may be also generated by

strong influence of shear forces or by reinforcement slippage.

For members having low shear span ratio ( , where l is the distance between maximum and

null moment sections and d is the effective depth of the section), say below 2.0-2.5, high shear

stresses have to be transferred across plastic zones and the member stiffness may become

reduced due to shear deformations; these can assume particular importance when cracks are

l d⁄

Page 43: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 19

open over the entire depth of the section, causing shear forces to be transferred by aggregate

interlock (a rather brittle mechanism and prone to sliding deformations) and by dowel action

(which may severely affect the steel-concrete bond in the vicinity of cracks).

Reinforcement slippage inside the concrete core, due to bond deterioration, is also responsible

for pinched diagrams, particularly for beam-column joints. Slippage of tensioned steel bars

induce higher deformations taking place without (or with low) increase of resistance due to

lack of bond mechanism, i.e. at very low (or null) stiffness; therefore, only after the slippage

deformations of the previous cycle have been sufficiently recovered to produce crack closure,

the stiffness may increase again due to re-activation of the compressive stresses on concrete.

The described features are the most relevant ones as shown by experimental evidence and have

constituted the basis for development of the so-called phenomenological analysis models for

numerical simulation of member behaviour. In these kind of models the observed response fea-

tures, viz:

• the unloading stiffness degradation,

• the reloading stiffness deterioration, due to both the Bauschinger effect and the open cracks,

• the pinching effect, resulting from crack closure and related with reinforcement asymmetry,

high shear forces or reinforcement pull-out and slippage and

• the strength deterioration caused by concrete degradation in compression or buckling of

reinforcement,

are approximately simulated by means of a set of rules relying upon a skeleton, or basic, curve

(typically the monotonic envelope or even others based on that one). These rules are usually

set up in an empirical basis and calibrated with experimental results.

2.3 Numerical modelling strategies

2.3.1 General

Numerical analysis of reinforced concrete building structures under seismic loading can be

performed, as any other type of structure, by means of the Finite Element Method. An ade-

quate response assessment requires realistic non-linear behaviour modelling of the constituent

material, which, for reinforced concrete, is better accomplished by adopting separate models

for concrete, steel and, eventually, their interaction through bond.

Page 44: Seismic Assessment of Reinforced Concrete Frame Structures

20 Chapter 2

In the most general three-dimensional analysis, the behaviour of structural members is

described by integration of the stress-strain relations of the constituent materials at the control

point level (typically the Gauss points). However, despite the computer power available nowa-

days and the large progress made to date in the field of constitutive modelling of materials, this

refined analysis may easily become prohibitive for multi-storey structures, particularly if time

step dynamic calculations have to be performed for seismic assessment purposes. Indeed, time

domain seismic analysis often requires several calculations for different input motions in order

to draw reliable conclusions about the seismic behaviour, which means that computational

demands may increase exponentially. Therefore, this type of detailed modelling has been

restricted to the analysis of individual members or sub-assemblages, mainly to have a better

understanding of some local phenomena rather than the global response.

On the other hand, the specific behaviour of building structures under horizontal actions (those

of major concern in the present work), for which the horizontal degrees-of-freedom play an

important role, permits the idealization of simplified structural models for the assessment of

global response parameters.

Typically, relatively simple models consisting of few degrees-of-freedom are used to simulate

the behaviour of groups of structural members, or even the entire structure. That is the case of

the so-called storey-models, as the shear-beam model, consisting of one horizontal displace-

ment per storey and having the behaviour governed by relations between the storey shear force

and the corresponding inter-storey drift (which can also be used for spatial analysis by adopt-

ing two horizontal translations and one rotation). Such type of models can be used when the

inelastic behaviour source is mainly located in columns, typically in the so-called weak-col-

umn strong-beam frame systems. The inter-storey behaviour is simulated by non-linear phe-

nomenological models where cracking and yielding are included along with several hysteretic

rules taking into account the main features referred in 2.2. In the limit, an even simpler model,

involving only a single degree-of-freedom in each horizontal direction and ruled by base shear

- top displacement relationships, can be adopted for structures fairly controlled by its first

mode of vibration.

However, the use of such simplified models is restricted to regular structures, both in stiffness

and mass distribution terms, and only global response features can be obtained, viz time histo-

ries and peak values of horizontal displacements or base shear. Further detail concerning ine-

Page 45: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 21

lastic effects and damage in individual members, obviously cannot be predicted.

In between these two extreme types of modelling strategies, the so-called member-type models

appear as the best available compromise of structural response detail and efficiency/simplicity.

A single element in the model is used for each member in the structure (a beam, a column, the

inter-storey portion of a shear-wall, etc.), and the behaviour is most often described by phe-

nomenological models, although stress-strain constitutive relationships at the material level

can be also used along with some member type modelling.

Although not accounting for minor details of geometry and exact reinforcement arrangement,

member models provide a description of the seismic damage and non-linear effects along the

member length, the accuracy and detail of which being dependent of the powerfulness of the

underlying formulation.

Member type models are worldwide used for non-linear dynamic analysis of reinforced con-

crete building structures, because computational requirements are quite manageable (even for

three-dimensional analysis) in view of the outcome they can provide. For this reason, and since

the model developed in this work fits in this modelling strategy, member models are further

discussed next in order to establish the framework of the present developments.

2.3.2 Member type models

The simplest and earliest member models have been first developed for uniaxial bending dom-

inated elements, where shear forces are not of major importance. This appears quite reasonable

because, on one hand the flexural behaviour has been always better understood than the shear

one, and, on the other hand, the earthquake-resistant design trends emerged in the past 25

years, being based on the weak-beam strong-column philosophy, lead to relatively slender

beams mainly controlled by uniaxial bending rather than shear.

However, other structural members, viz the first floor columns where inelastic behaviour is

likely to occur (even for the weak-beam strong-column type of structures), involve interaction

of axial load and bending (possibly biaxial) which may be important to incorporate in the

model. Soon this aspect has attracted the researchers attention and some models were proposed

over the past two decades to take it into account, particularly if significant fluctuation of axial

load is likely to occur during the seismic response. Similarly, the bending-shear interaction

Page 46: Seismic Assessment of Reinforced Concrete Frame Structures

22 Chapter 2

problem has been investigated, although fewer research work can be found due to the problem

complexity in the seismic analysis context.

In the present work, emphasis is put in uniaxial bending behaviour where the presence of static

axial force (due to gravity loads) is taken into account, but not its variation during the

response; the complete interaction problem is thoroughly addressed and the most significant

models are reviewed in Fardis (1991). Furthermore, any inelastic shear behaviour interaction is

not considered herein.

Member modelling comprises two main issues:

• the element model, where assumptions are made concerning the stiffness distribution and

the force or displacement fields along the member, which allow more or less simplified for-

mulations;

• a hysteretic behaviour model, defining the generalized force-displacement relationships to

be adopted in the inelastic zones of the member.

In the context of building structure response to horizontal actions, particularly seismic ones,

the inelastic behaviour is typically located at and near the member ends, where bending

moments are maximum, and the inelastic zone development is directly related to the moment

distribution along the member.

Thus, two basic models proposed in the 1960s have set up the framework for member model-

ling, both assuming the inelastic behaviour lumped in “point hinges” at member ends:

• The so-called two-component model proposed by Clough et al. (1965) consists of two beam

elements associated in parallel as shown in Figure 2.3-a). One is an elastic-perfectly-plastic

beam and accounts for the elastic behaviour: prior to yielding at one of the member ends, it

behaves elastically, the stiffness matrix being that of the classical elastic beam; after the

onset of yielding at one end, the tangent stiffness matrix contribution is that of a beam with

a hinge at that end and, when yielding occurs also at the other end, the stiffness contribution

becomes that of a beam with hinges at both ends. The other element behaves continuously

elastic, having the stiffness equal to that assumed in the post-yielding range ( , where k is

the elastic stiffness and p is the fraction attributed to hardening), which means that the first

component has to account for the remaining elastic stiffness prior to yielding.

The superposition of diagrams yields an elastoplastic behaviour and, because the first com-

pk

1 p–( )k

Page 47: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 23

ponent unloads/reloads parallel to the elastic branch and the second component remains

elastic, the resulting hysteretic behaviour is bilinear. Due to the parallel association of com-

ponents, this model is a stiffness based one and stiffness matrix contributions can be

directly obtained and added together.

Figure 2.3 Member modelling: a) two component model (Clough et al. (1965)) and b) one com-

ponent model (Giberson (1967))

• The one-component model, as introduced by Giberson (1967) and shown in Figure 2.3-b),

is an association in series of one elastic element with two non-linear rotation springs at each

end where the inelasticity is lumped; the former accounts for elastic deformations and the

latter are governed by a non-linear behaviour model to account for inelastic rotation compo-

nents. The fact that point hinges (zero length) are considered leads to inelastic rotation at a

given hinge uniquely determined as a function of the moment at that hinge (no coupling of

deformations) and the corresponding flexibility matrix becomes diagonal. By assuming the

inflection point at mid-span, the equivalence between each half member and a cantilever

beam allows the inelastic rotation to be estimated by curvature integration and lumped into

each hinge; hence, the moment-rotation relationship can be derived and used for the hinge

state determination. According to the association in series, the inelastic flexibility matrix

associated with the non-linear springs is then added to the flexibility matrix of the linear

elastic element.

MA MB

θA θB1 p–( )k

pk

Elastic sections turning to plastic for θ θy>

θA

θB

θAe

θBe

M

θ

M

θ

1 p–( )k

pk

1

M

θk

1

+ =

θAp

θBp

θA θBIF

Elasticnon-linearrotational springs

a) Two-component model

b) One-component model

Page 48: Seismic Assessment of Reinforced Concrete Frame Structures

24 Chapter 2

The two component model, although conceptually very simple, has the major drawback of not

being able to simulate stiffness degradation; instead, in the one-component model this problem

is overcome by the use of a hysteretic behaviour model at each hinge, incorporating rules for

that and other features described in 2.2. However, the one-component model is not problem-

free: the lumped inelasticity is just an approximation of the real spreading of the inelastic zone

as the moment distribution changes, and the moment-rotation relationships are derived under

the assumption of inflection point at mid-span which is not satisfied in many cases (e.g. col-

umns, beams with gravity loads or members with different contents of top and bottom rein-

forcement).

Despite these limitations, the uncoupling between inelastic rotations of member-ends and the

possibility of using any hysteretic model have rendered the one-component model rather popu-

lar and widely used. Additionally, its simplicity, low computational cost and numerical stabil-

ity and robustness have motivated the development of several hysteretic models used for both

local (moment-curvature) and global (force-displacement or moment-rotation) behaviour anal-

ysis.

Some of the most representative phenomenological models developed to date are briefly

referred next, following the chronological order:

• The Clough and Johnston model (Clough and Johnston (1966)) is based on a bilinear basic

curve with unloading parallel to initial stiffness, but including degradation of reloading

stiffness by pointing to the maximum deformation previously reached.

• Anagnostopoulos (1972) has complemented the previous model with the ability of degrad-

ing unloading stiffness according to the maximum reached deformation.

• The Takeda model (Takeda et al. (1970)) is based on a trilinear basic envelope, accounting

for elastic, cracked and hardening stages, which may be different in the two loading direc-

tions. It includes unloading stiffness degradation after the onset of cracking (thus account-

ing for dissipation already at the cracked stage) according to the maximum reached

deformation, as well as the same rule for reloading stiffness deterioration as for the previous

models. Small amplitude cycles are taken into account but neither pinching nor strength

deterioration are considered. Although requiring a wide set of hysteretic rules (16), this

model has become a reference milestone of phenomenological models.

• Otani (1974) and Litton (1975) have proposed simplifications of the Takeda model, by

using bilinear basic curves and a smaller set of slightly modified hysteretic rules.

Page 49: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 25

• The Banon model (Banon (1980)) is also based on a bilinear curve, with hysteretic rules

similar to those of the Takeda model but already including the modelling of the pinching

effect.

• Roufaiel and Meyer have developed another model (Meyer et al. (1983), Roufaiel and

Meyer (1987)) which includes the features of the previous one and also takes into account

the strength deterioration due to cyclic effect beyond a certain critical deformation (related

with a critical compressive strain in concrete).

• Costa and Campos Costa have proposed a model (Costa and Campos Costa (1987), Costa

(1989)) which considers a trilinear basic curve (different in positive and negative direc-

tions), along with hysteresis rules similar to the Takeda model (slightly modified), including

the pinching effect and the strength degradation. Before yielding the model behaves biline-

arly without stiffness and strength reduction and small cycles are taken into account by

monitoring relative maximum deformations.

• The Park et al. (1987b) model basically includes the feature of the previous one, although

with slightly different rules: unloading stiffness degradation is achieved by using a common

target point for unloading branches and is activated after cracking occurrence; strength deg-

radation is taken into account with the concept of damage contribution from dissipated

energy as introduced by Park et al. (1984).

This type of models arise from the knowledge of experimentally observed phenomena and typ-

ically reproduce the stiffness variations during the loading process by means of piecewise lin-

ear type curves. Hysteretic rules are usually based on the history of some response parameters,

allowing the state determination to be performed without additional effort.

Other type of models provide continuous hysteretic relations, of which the Ramberg-Osgood

model is an example; it has been derived from the steel stress-strain constitutive law intro-

duced by Ramberg and Osgood (1943) and adapted by Jennings (1963) for the hysteretic

behaviour of reinforced concrete members.

Additionally, it is worth referring the so-called rate-type or differential constitutive models,

derived from the endochronic theory formulation in Ozdemir (1981) and Brancaleoni et al.

(1983), which is formally identical to viscoelasticity but where time is replaced by a deforma-

tion measure called “intrinsic time”. Examples are the Bouc-Wen model (Wen (1980)), widely

used in non-linear stochastic analysis, and the Spacone model (Spacone et al. (1992)), where

Page 50: Seismic Assessment of Reinforced Concrete Frame Structures

26 Chapter 2

the tangent stiffness is approximated by a continuous function of the current state and of model

parameters; this means that a non-linear incremental constitutive law has to be previously

obtained from integration of the differential relation (expressing the tangent stiffness) within

each load step. In the context of usual stiffness based finite element programs, this constitutes

a major drawback, but, as outlined by Spacone et al. (1992), it fits quite naturally in the flexi-

bility based algorithm described later in this chapter; nevertheless, the large set of required

model parameters to be selected, still represents a major problem in the adoption of such type

of models.

Most of the above referred hysteretic models have been largely used along with the one-com-

ponent model due to its versatility and simplicity. However, the already mentioned limitations

inherent in such member model, particularly those related with the difficulty of defining

moment-rotation laws and the inconsistency arising from the assumption of a fixed inflection

point at mid-span, led to the development of distributed inelasticity member models in order to

more closely follow the member behaviour coherently with its actual characteristics and load-

ing conditions.

2.3.3 Distributed inelasticity member models

The generalization of the one-component model to account for inelasticity spread along the

member was first introduced by Otani (1974). In his proposal, the non-linear rotational springs

for member inelastic behaviour are replaced by two inelastic finite length elements. However,

two other non-linear rotational springs are additionally considered (with zero length) to repre-

sent the fixed-end rotations at the beam-column interface due to reinforcement slippage inside

the concrete core of the joint. Upon recognition of the inconsistency of the one-component

model resulting from considering a fixed inflection point at mid-span when applied to generic

linear moment distributions, Otani has derived the inelastic flexibility matrix associated with

the two non-linear elements as a function of the current location of the inflection point. Each

element part at each side of the inflection point is identified with an equivalent cantilever,

where the free-end displacement and the rotation - fixed-end moment relations are prescribed.

However, by adopting some simplifying approximations, the approach results in a non-sym-

metric flexibility matrix, unless the further assumption of lumping inelasticity at member-ends

is introduced. Therefore, the proposal ends up in the lumped inelasticity group, although the

inelastic stiffness takes into account the variation of location of the inflection point. Despite

Page 51: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 27

the non-significant final outcome of the model concerning spread of inelasticity, it was the first

to recognize the importance of fixed-end rotations in predicting the seismic response of rein-

forced-concrete structures.

Soleimani et al. have proposed the first model actually taking into account the gradual spread

of inelasticity along the member length as a function of the loading history (Soleimani et al.

(1979)). Inelastic zones gradually increase from the beam-column interface, while the rest of

the member remains elastic. The behaviour of inelastic zones is controlled by a moment-curva-

ture relationship of Clough type (Clough and Johnston (1966)) prescribed at the member end

sections, where point hinges are also considered to account for fixed-end rotations at the beam-

column interface. These are related to curvature at the corresponding end section by means of

an “effective length” factor assumed constant during the response.

Several other authors have proposed slightly different models along the same trend-line

(Arzoumanidis and Meyer (1981), Roufaiel and Meyer (1983), Roufaiel and Meyer (1987)),

basically to introduce refinements in the hysteretic model, viz the pinching effect accounting

for shear interaction, the strength degradation and the (constant) axial force interaction with

the basic curve of the model.

The Soleimani proposal was further elaborated by Filippou and Issa (1988), and cast in a more

complete context. The member is subdivided into different sub-elements, each accounting

exclusively for a single effect, namely: the elastic behaviour, assumed cracked before yielding

of reinforcement, the inelastic behaviour due to bending (either concentrated in point hinges or

spreading along the member) and the fixed-end rotations at the beam-column interface. Sub-

elements are associated in series, for which the relevant contributions to the flexibility matrix

are summed-up; the same applies for contributions of member-end rotations arising from each

sub-element.

The point hinge idealizations used in this model (for concentrated inelastic bending or fixed-

end rotations) are based on a bilinear moment-rotation envelope with constant post-yielding

stiffness. For inelastic bending, that stiffness is an approximation of the non-linear post-yield-

ing relation, calculated for a pre-defined ultimate moment capacity, and the hysteretic behav-

iour follows a Clough type model. For the fixed-end rotation, the moment-rotation relation is

derived using the detailed model proposed by Filippou et al. (1983), according to which a

Page 52: Seismic Assessment of Reinforced Concrete Frame Structures

28 Chapter 2

beam-column joint model, representing a particular connection under analysis, is subjected to

monotonic increasing beam end moments to give rise to the concentrated rotations due to bar

pull-out at the beam-column interface.

The spread of inelastic bending is considered along member end zones, with non-decreasing

lengths, where the yielding moment is exceeded. An average stiffness is assumed in these

zones entirely determined by the corresponding end section behaviour, which is governed by a

modified Clough type model, and a special algorithm is used to control the advancement of

inelastic zones.

This model was further improved in a subsequent study by Filippou et al. (1992) to include

another sub-element accounting for shear distortions in inelastic zones and shear sliding at the

beam-column interface. Additionally, constant axial force - bending moment interaction was

also included in the basic curve of the model.

Takayanagi and Schnobrich (1976,1979) have proposed another type of member model, by

dividing the element into several short sub-elements (finite length springs) along the member

axis, each of them governed by a non-linear moment-rotation (or curvature) model. Properties

(namely stiffness) are assumed constant along each sub-element length which is controlled by

the moment at its mid-point. The axial force - bending moment interaction is taken into

account by a limit surface for each spring and static condensation is used to lump the element

behaviour into the element end springs; this model fits in the so-called multi-slice model type.

However, it may suffer from the problem of unbalanced forces developing in internal sub-ele-

ments whose degrees-of-freedom are not explicitly controlled in the global non-linear algo-

rithm scheme; indeed, since the elimination of residual forces does not take them into account,

the internal resisting moments calculated via deformations at each sub-element do not neces-

sarily match the applied moments obtained via equilibrium conditions; consequently, equilib-

rium is locally violated, which often introduces numerical instability.

This model, as well as others with distributed inelasticity (Hellesland and Scordelis (1981) ,

Mari and Scordelis (1984)), are typically based on the classical stiffness method using the well

known cubic Hermitian polynomials to approximate displacements along the element. Both

the element stiffness matrix and the nodal equivalent forces are obtained by integration of the

section stiffness and force distributions, respectively, duly weighted by deformation shape

Page 53: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 29

functions derived from the Hermitian polynomials, as for the linear elastic elements.

However, cubic polynomials give the exact solution for prismatic members having uniform

linear elastic properties and without loads applied in the span (thus, only linearly distributed

moments originating linear distribution of curvatures). Once the non-linear behaviour becomes

significant, the curvature distribution may fairly deviate from the linear one, which means that

the cubic representation of displacements is no longer adequate and numerical problems are

very likely to appear. In the general context of the Finite Element Method, the problem is over-

come by mesh refinement in the potential inelastic regions, but this does not fit with the scope

of member modelling.

Some proposals have been made with improved representation of internal deformations.

Menegotto and Pinto (1977) proposed to achieve this improvement by combined approxima-

tion of both the section deformations and flexibilities, the latter being assumed linearly distrib-

uted between a few controlled sections (thus, equivalent to hyperbolic stiffness distribution).

The recognition of the non-adequacy of cubic polynomials to approximate inelastic member

deformations led to the improvement of displacement interpolation by introducing the so-

called variable interpolation functions. These were proposed first by Mahasuverachai (1982)

for piping and tubular structures, and adapted to reinforced concrete members by Kaba and

Mahin (1984) along with section layer discretization. Typically, these functions are derived

from force interpolation polynomials (obtained from equilibrium conditions, independently of

the element state) corrected by the current flexibility distribution and the flexibility matrix of

the member. A mixed approach is used, where both deformation and force interpolation func-

tions are adopted. However, the model is found to contain inconsistencies (Taucer et al.

(1991)) that give rise to numerical problems; particularly, the state determination is such that

equilibrium between section resisting forces (along the member) and section applied forces is

not satisfied.

The proposal is further developed by Zeris (1986) and Zeris and Mahin (1988), namely by

extending it to biaxial bending through fibre modelling and by improving the element state

determination process. The model succeeded to overcome the numerical problems found with

the Kaba and Mahin proposal and, particularly, equilibrium is guaranteed by a specific iterative

procedure. The response of elements with softening behaviour that could not be analysed with

Page 54: Seismic Assessment of Reinforced Concrete Frame Structures

30 Chapter 2

previous models, was finally simulated successfully. The proposed model has shown satisfac-

tory performance (Taucer et al. (1991)) but the element state determination was found not very

clear, surely due to its derivation from ad-hoc corrections of the Kaba and Mahin model. Nev-

ertheless, these two last proposals, by involving the use of force interpolation functions, have

pointed-out the promising framework of the flexibility approach.

The formulation of non-linear flexibility based frame elements has been cast in the form of a

unified and general theory (Taucer et al. (1991), Spacone et al. (1992) and Spacone (1994))

derived from the framework of mixed finite element methods (Zienkiewicz and Taylor (1989)).

Use is made of the two field mixed method, in order to address the element state determination

(indeed the most critical task in the flexibility based formulation) and to clarify its strong con-

nection with numerical implementation. Thus, a consistent and elegant element state determi-

nation scheme is proposed, whose insertion in classical stiffness based finite element programs

appears rather straightforward. Moreover, the formulation is suitable to accommodate any kind

of section constitutive relationship, either a global one (e.g. moment-curvature at the section

level) of phenomenological or differential type, or an implicit one arising from integration of

the local stress-strain behaviour at the fibre level.

The general formulation for flexibility based elements typically requires a few control sections

distributed along the element whose behaviour is monitored. Only force interpolation func-

tions are used, because they have the major advantage of being exact regardless of the dam-

aged state of the member. The element flexibility matrix is obtained from integration of the

flexibility distribution (known at the control sections) and an internal iterative scheme is used

to obtain the element resisting forces for imposed displacements in the element nodes; such

scheme is driven by residual displacements to be gradually eliminated while strictly preserving

equilibrium along the element.

The formulation has been associated with fibre discretization of the section by Taucer et al.

(1991) to develop a general three-dimensional beam-column element where biaxial bending

with axial force interaction is automatically taken into account. In a parallel work by Spacone

et al. (1992), a differential constitutive relation is adopted for uniaxial bending, taking into

account strength and stiffness degradation and pinching through a measure of accumulated

damage; since the interaction of varying axial force is not considered, the model is mainly

appropriate for beams or columns with low axial load level. However, that includes also the

Page 55: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 31

fixed-end rotations modelling by means of rotational point hinges at the member ends. The

most remarkable aspect is the natural and elegant way how the proposed element state determi-

nation scheme can accommodate the association in series of two or more elements behaving

non-linearly; this problem had been already addressed in a previous work by Filippou and Issa

(1988), but the solution proposed there was more cumbersome and tricky. Therefore, in our

opinion, Spacone proposal appears more clear in dealing with the association in series of any

number of non-linear elements and rather suitable to incorporate the wide set of sub-elements

proposed in Filippou et al. (1992) for separate simulation of each non-linear behaviour source.

For its importance in the present work context, the general formulation of flexibility based ele-

ments is further discussed in 2.4, following the basic steps of previous works (Taucer et al.

(1991), Spacone et al. (1992)).

Finally, although not strictly fitting in the member model family, it is worth referring another

recent modelling strategy found in Izzuddin (1991), Izzuddin et al. (1994) and Karayannis et

al. (1994). It follows a displacement based formulation, and classifies the non-linearity sources

in two basic types: one for the effects of concrete cracking and the non-linear compressive

response of concrete, and another for the effects of open cracks, concrete crushing (and con-

finement) and the post-yielding behaviour of steel. The first source, although non-linear, is

assigned an elastic behaviour, but a higher order element is proposed whose formulation is

based on a quartic shape function for transverse displacements. The second source is modelled

by a classical cubic Hermitian element, for which the non-linear behaviour is controlled at two

Gauss points by means of a layer approach, separately accounting for the behaviour of steel

and of unconfined and confined concrete. The analysis is performed by an adaptive procedure:

all elements start behaving elastically and modelled by a single elastic quartic element; as ine-

lasticity takes place in critical regions, inelastic cubic layered elements are gradually inserted

as needed. Automatic mesh refinement techniques are required, with particular care on setting

up initial conditions for new elements when they are inserted; however, the model appears to

be quite efficient and accurate when compared to results from discretizations with inelastic

layered elements inserted at the beginning of the analysis.

The above review is meant to cover, in a historical perspective, the currently available range of

member models for uniaxial bending conditions. Several others can be found in the literature

to account for varying axial load interaction with bending (eventually biaxial), which were not

Page 56: Seismic Assessment of Reinforced Concrete Frame Structures

32 Chapter 2

referred herein since that is not relevant for the developments envisaged in the present work.

This problem was investigated by Coelho (1992) through the analysis of some frame structures

where the column axial load variation was approximately taken into account. Despite some-

what significant fluctuations of axial loads in external columns (due to overturning moments),

the results in terms of maximum storey displacements and ductility demands were just slightly

affected, showing some trend for lower values comparatively to the assumption of constant

axial loads. Thus, the overall conclusion pointed to a not very significant influence of that

problem on the global response. For this reason, and because the consideration of that issue in

the model developed herein would lead to extremely complicated numerical schemes, it has

been decided to restrict the present stage of development to the interaction of static axial force

- bending moment, i.e. just on the skeleton curve of the adopted phenomenological model.

The review has culminated with the formulation of flexibility based elements, which, in our

opinion, appears as the most promising and elegant formulation developed to date. To a certain

extent, this formulation brings back to evidence the important role the force method plays in

structural analysis (particularly of frame structures), that was somewhat “forgotten” due to the

advent and massive application of the displacement based Finite Element Methods. Both tech-

niques are clearly shown to be compatible (Spacone et al. (1992)), thus opening the doors to

further developments of complementary tools for structural modelling.

Indeed, it appears quite convenient that, for an integrated discretization of a complete struc-

ture, several types of elements can be mixed-up in order to better simulate the behaviour of

each constituent. As a typical example, for the complete numerical model of a building struc-

ture it is desirable to consider adequate plate or shell elements to simulate slabs, plane ele-

ments to model walls, member models for beams and columns, joint elements for beam-

column joints (of which a very interesting proposal is reported in Monti et al. (1993)), founda-

tion elements to include soil-structure interaction and, finally, to make them all compatible in a

unique analysis model. Fortunately, that is the trend of general purpose computer codes emerg-

ing over the last decade, an example of which is CASTEM2000 (CEA (1990)) where the ele-

ment model developed in the present work has been implemented (further references to the

code are included in Chapter 4).

Page 57: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 33

2.4 General flexibility formulation for beam-column elements

For a comprehensive perspective of the new approach developed in the present work, the gen-

eral flexibility formulation for beam-column elements is reviewed in the following sections.

The description is mainly based on previous work (Taucer et al. (1991), Spacone et al. (1992),

Spacone (1994)) where further details can be readily found. Slightly different conventions and

notation are adopted as introduced in 2.4.1 and rigid zones are considered at the element ends,

called rigid lengths for briefness. Furthermore, an attempt is made in 2.4.2 to clearly distin-

guish all the involved variable spaces, ranging from the global structure to the local section

levels.

The most relevant topics of the theoretical background for the flexibility formulation are

recalled in 2.4.3 in contrast with the classical stiffness method, highlighting the difficulties

associated with the element state determination in the flexibility context arising from the non-

availability of displacement shape functions.

Special attention is given to the element state determination in 2.4.4, as the most delicate task

in the flexibility formulation, and, after the presentation of the algorithm to handle only nodal

loads, the process to include also element applied loads is described as an extension and adap-

tation of the previous one. However, the presentation is made without restricting references,

neither to the type of section model (fibre or global) nor to the type of element applied loads, in

order to keep its generality.

Finally, it is worth recalling that this formulation can be also included in the context of classi-

cal stiffness based finite element computer codes, despite its somewhat higher implementation

cost. Indeed, given an input of nodal displacement increment for each element, the algorithm

provides the usual output of the state determination of any classical finite element, i.e., the cor-

responding increment of element restoring forces and the updated stiffness matrix, ready for

assembly in the global structure restoring force vector and stiffness matrix, respectively.

2.4.1 Conventions and notation

Figures 2.4 and 2.5 show the three different reference frames used in the present study: the glo-

bal coordinate system (X,Y,Z), the element local system (x,y,z) and the section local reference

system (xs ,ys ,zs).

Page 58: Seismic Assessment of Reinforced Concrete Frame Structures

34 Chapter 2

Associated with these axis systems, the force, displacement and generalized stress and strain

vectors are written according to the following general rule: boldface upper case letters denote

force and generalized stress vectors (for which Q and S are the reserved letters), while bold-

face lower case letters are used for displacement and generalized strain vectors (denoted by u

and e).

Figure 2.4 Spaces of variables and axis systems: from the global to the local element level

Figure 2.4 shows all the variable spaces used in the analysis at global and element level,

whereas Figure 2.5 refers to the variables at section level. Translation and rotation displace-

ment components are denoted by uξ and θξ, respectively, where ξ stands for the associated axis;

similar notation, Fξ and Μξ, is used for the corresponding force and moment components. The

Z

Y

X

zy

x

1

2

1

2

x

E1

E2x

E1

E2

Ll1 l2

FZG

MZG

FYG

MYG

FXG

MXG

QG uG( , )QE uE( , )

a) Global Reference System andSpace of Global Nodal Variables

b) Element Global Space of Variables

Q1E

Q2E

Qe ue( , )

c) Element Local Variables

Q1e

Q2e

zy

Q1f

Q2f

Qf uf( , )

d) Flexible Element Local Variables(with rigid body modes)

Q u( , )

e) Flexible Element Local Variables(without rigid body modes)

zy

Page 59: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 35

positive direction of each component is the same of the corresponding axis. Superscripts are

used to identify the variable space and subscripts refer to a specific element node (for the total

element, i.e. with rigid lengths included) or an end section (for the flexible element part).

Figure 2.5 Space of variables at the local section level.

According to Figure 2.4-a), the force and displacement vectors, at the global level are written

as follows

(2.1)

while at the element level with rigid body modes, the following notation is adopted

(2.2)

The superscript η in Eq. (2.2) holds for E, e and f, identifying the complete variable spaces of

the element, viz that with axes parallel to the global system (Figure 2.4-b)), the one referring to

the local system (Figure 2.4-c)) and the one of the flexible element part (Figure 2.4-d)).

The basic system without rigid body modes is characterized by the degrees of freedom (d.o.f.)

shown in Figure 2.4-e), and the corresponding force and displacement vectors are defined by

(2.3)

At the section level, the generalized stress and strain and the displacement vectors are written

in the local axis system (Figure 2.5) and given by

zy

xE1 xs

zs ysMz

My

Mx

Vz Vy

Nxns

QGFX

G FYG FZ

G MXG MY

G MZG

T= and uG

uXG uY

G uZG θX

G θYG θZ

GT

=

Qη Q1η

Q2η⎩ ⎭

⎨ ⎬⎧ ⎫

= and uη u1η

u2η⎩ ⎭

⎨ ⎬⎧ ⎫

= with

Qiη

FXi

η FYi

η FZi

η MXi

η MYi

η MZi

ηT

= uiη

uXi

η uYi

η uZi

η θXi

η θYi

η θZi

η

i 1 2,=( )

T=

Q Fx1Mx1

My1Mz1

My2Mz2

T= and u ux1

θx1θy1

θz1θy2

θz2

T=

Page 60: Seismic Assessment of Reinforced Concrete Frame Structures

36 Chapter 2

(2.4)

The local system (xs ,ys ,zs) is considered with the xs co-linear with the element x axis, but the

remaining local axes may be oriented in any other direction; if this is the case, then a local

orthogonal transformation must be performed over the relevant vectors in order to conform

with the element (x,y,z) reference system. The positive unit normal ns defines the local system

positive directions and, consequently, the positive directions for stresses, strains and displace-

ments. However, in this section and for the sake of simplicity, these two reference systems are

considered parallel and the subscript s is suppressed.

The usual meaning of the generalized stresses and strains is adopted, namely: axial force (N)

and strain (ε), transverse forces (V) and distortion rotations (β), twisting and bending moments

(M) and associated curvatures (ϕ). Shear distortions (β) are included although they depend on

the adopted beam formulation (e.g., in the Bernoulli case those distortions are neglected).

The externally applied actions can be the following:

• nodal actions, included in the and vectors, and, therefore, consisting of imposed

forces or displacements;

• element loads, usually forces and moments imposed along the element, either distributed or

concentrated, as schematically shown in Figure 2.6; in the present study, element loads are

considered only in the flexible part, since this corresponds to the most common situation.

Figure 2.6 Element applied loads

S S x( ) NxsVys

VzsMxs

MysMzs

T= =

e e x( ) εxsβys

βzsϕxs

ϕysϕzs

T= =

a a x( ) uxsuys

uzsθxs

θysθzs

T= =

QG uG

zy

xE1 E2

Lx

Pz

µz

Py

µy

Px µxpz

py

px

mz

my

mx

ns

Page 61: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 37

The vectors of distributed and concentrated element loads are given, respectively, by

(2.5)

and, in the most general case, the load distribution can be of any type and applied wherever

desired.

However, moments applied in the element are rarely considered for practical situations; hence,

only the following reduced vectors are used in this study

(2.6)

For equilibrium purposes (discussed in the following sections), force and moment resultants of

the element loads have to be defined. Between the flexible end section E1 and a generic section

with abscissa x, the vector of resultants (denoted by for force components and by for

moment ones) includes the contribution of all the loads applied in that zone and is written as

(2.7)

where moments are taken relative to the generic section. In particular, for the whole flexible

element, the total resultant vector is defined as

(2.8)

2.4.2 Relations between spaces of variables

The above mentioned spaces of variables are inter-related through topologic, geometric, equi-

librium and compatibility type conditions.

The topologic conditions relate the global nodal space with the element global space

and are stated by means of well known procedures of matrix structural analysis.

p p x( ) px py pz mx my mz

T= =

P P x( ) Px Py Pz µx µy µz

T= =

p p x( ) px py pz

T= =

P P x( ) Px Py Pz

T= =

R R x( ) x y z x y z

T= =

RL R L( )xL

yL

zL

xL

yL

zL

T= =

QG uG( , )

QE uE( , )

Page 62: Seismic Assessment of Reinforced Concrete Frame Structures

38 Chapter 2

The geometric conditions consist on axis system rotations and define the relation between the

global and the local spaces of variables at the element level; as already men-

tioned, this type of relation may have to be also applied at the section level if non-coincident

reference systems are used.

The transformation between the element local space and the flexible element space

is derived from equilibrium conditions of the rigid lengths and can be written as

(2.9)

with the transformation matrix given by

(2.10)

where, according to Figure 2.4-b), l1 and l2 stand for the element rigid lengths and I is the

(3x3) identity matrix.

Element rigid body modes are fixed by imposing the appropriate boundary conditions shown

in Figure 2.4-e) and the flexible element behaviour can be described by the reduced space of

variables . The equilibrium conditions, involving the fixing support reactions, the

forces and the total resultants of element loads , allow the following relation to be obtained

(2.11)

where the transformation matrix and the vector of element loads contribution , are

given, respectively, by

QE uE( , ) Qe ue( , )

Qe ue( , )

Qf uf( , )

Qe Tr Qf⋅= and uf TrT

ue⋅=

Tr

Tr

I[ ] 0[ ] 0[ ] 0[ ]

T1r[ ] I[ ] 0[ ] 0[ ]

0[ ] 0[ ] I[ ] 0[ ]

0[ ] 0[ ] T2r[ ] I[ ]

= with

T1r

0 0 00 0 l1–

0 l1 0

=

T2r

0 0 00 0 l2

0 l– 2 0

=

Q u( , ) Q

RL

Qf Tb Q⋅ Qpf+=

Tb Qpf

Page 63: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 39

(2.12)

Assuming small deflections, the relation between displacement vectors with and without rigid

body modes is given by

(2.13)

Following a very similar procedure, the relation between the element forces Q and the generic

section stress can be obtained. Actually, writing the equilibrium equations for the portion

of element between the end section E1 and the generic section at abscissa x, including the

resultants of element loads applied up to x, leads to the following relation

(2.14)

where the matrix and the vector (with the contribution of the element applied

loads) are given by

(2.15)

In turn, the compatibility relation between displacements u and the section deformations

is given by the following integral form

Tb

1 0 0 0 0 00 0 0 1 L⁄ 0 1 L⁄0 0 1 L⁄– 0 1 L⁄– 00 1 0 0 0 00 0 1 0 0 00 0 0 1 0 01– 0 0 0 0 0

0 0 0 1 L⁄– 0 1 L⁄–0 0 1 L⁄ 0 1 L⁄ 00 1– 0 0 0 00 0 0 0 1 00 0 0 0 0 1

= and Qpf

0

zL L⁄

yL L⁄–

000

xL–

yL– z

L L⁄–

zL– y

L L⁄+

xL–

00⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

u TbT

uf⋅=

S x( )

S x( ) b x( ) Q⋅ Sp x( )+=

b x( ) Sp x( )

b x( )

1– 0 0 0 0 00 0 0 1 L⁄– 0 1 L⁄–0 0 1 L⁄ 0 1 L⁄ 00 1– 0 0 0 00 0 x L⁄ 1–( ) 0 x L⁄ 00 0 0 x L⁄ 1–( ) 0 x L⁄

= and Sp x( )

x–

zL L⁄– y–

yL L⁄ z–

x–

x yL L⁄( ) y–

x zL L⁄( ) z–⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

e x( )

Page 64: Seismic Assessment of Reinforced Concrete Frame Structures

40 Chapter 2

(2.16)

which can be derived from the virtual force principle.

2.4.3 Theoretical background

This section presents the theoretical bases for the general flexibility formulation of beam-col-

umn elements, written in the incremental form of non-linear analysis. The formulation

accounts for the presence of nodal loads only, since the case of element loads can be dealt with

equivalent nodal forces, conveniently taken into account as explained later.

At the section level, the force-deformation relation can be stated as desired: either by means of

a fibre discretization (Taucer et al. (1991)), each fibre being controlled by a stress-strain con-

stitutive material law, or by a global section constitutive relation in terms of generalized forces

and displacements (Spacone et al. (1992)).

The flexibility method is presented after a preliminary review of the classical stiffness method

in order to recall their main differences in the analysis procedure, particularly for non-linear

behaviour, and to identify the phases where adaptations are needed for integration of the flexi-

bility method in classical stiffness based finite element codes.

2.4.3.1 Stiffness method

The formulation based on the stiffness method requires the adoption of displacement shape

functions to approximate the kinematics inside the element in terms of member end displace-

ments with rigid body modes. Denoting by ∆ the increments of the relevant quantities, the

approximation of the displacement and deformation fields is done by

(2.17)

where is a differential operator, depending on the section formulation, is the matrix of

the displacement shape functions and contains the deformation shape functions. Eqs.

(2.17) represent the compatibility conditions between section deformations (and displace-

u bT x( ) e x( )⋅ xd0

L

∫=

∆a x( ) N x( ) ∆uf⋅=∆e x( ) ∂ ∆a x( )⋅= ⎭

⎬⎫

∆e x( ) B x( ) ∆uf⋅=⇒ with B x( ) ∂ N x( )⋅=

∂ N x( )

B x( )

Page 65: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 41

ments) and member end displacements .

Increments of section forces and deformations can be related by the local stiffness matrix

according to the linearized form

(2.18)

and the equilibrium between increments of element forces and section forces can

be stated with the principle of virtual displacements by

(2.19)

written upon an equilibrated distribution of forces. Substitution of Eq. (2.18) in Eq. (2.19),

which must hold for any choice of the virtual displacement vector , leads to the classical

stiffness relationship at the element level , where the stiffness matrix is

readily computed by

(2.20)

Eq. (2.19) can be also written to yield the element resisting forces

(2.21)

in equilibrium with a given distribution of section resisting forces .

The set of analytical expressions is now complete for the classical stiffness method. The non-

linear behaviour requires an iterative procedure at the global structural degrees of freedom, for

which the Newton-Rapshon scheme is most often adopted. The typical sequence of tasks is

schematically described in the flowchart of Figure 2.7 and develops as follows:

• for a given force increment at the structure level , or an iteration corrective incre-

ment, an estimate of displacement increment is obtained using a global stiffness

matrix (initial, tangent or other, depending on the selected variant of the Newton-Raphson

method);

∆uf

k x( )

∆S x( ) k x( ) ∆e x( )⋅=

∆Qf ∆S x( )

δufT

∆Qf⋅ δeT x( ) ∆S x( )⋅ xd0

L

∫=

δuf

∆Qf Kf ∆uf⋅=( )

Kf BT x( ) k x( ) B x( )⋅ ⋅ xd0

L

∫=

Qrf( )

Qrf BT x( ) Sr x( )⋅ xd

0

L

∫=

Sr x( )( )

∆QG( )

∆uG( )

Page 66: Seismic Assessment of Reinforced Concrete Frame Structures

42 Chapter 2

• increments of element displacements are obtained from the structural ones ,

by means of successive transformations according to 2.4.2, and the element state determina-

tion starts (as highlighted in Figure 2.7);

• section deformation increments are then computed by Eq. (2.17) and allow the

deformations to be updated;

• constitutive laws of sections are usually written in terms of , thus permitting to directly

obtain the local stiffness matrix and the section resisting forces , which are then

passed to the element level according to Eqs. (2.20) and (2.21); the element state determina-

tion is thus complete;

• the element resisting force vectors and the stiffness matrices are transferred to

the structure level (again according to relations given in 2.4.2), in order to check equilib-

rium against external loads and, if required, proceed with further corrections.

Figure 2.7 Main tasks of the classical stiffness based state determination process

These steps constitute the well known procedure at the core of the widely used stiffness based

methods of finite element analysis, for which further details are deemed unnecessary. How-

ever, as noted before, the deformation shape functions are generally derived for linear

elastic behaviour of prismatic members and, therefore, remain valid only for such conditions.

Thus, unless a more refined discretization is adopted for the member, the approximation

obtained with just one element per member in the non-linear range may become rather crude.

Typically, for monotonic analysis, this approximation leads to a solution stiffer than the “cor-

rect” one because the functions are not able to describe the more complex deformed

∆uf( ) ∆uG( )

∆e x( )( )

e x( )

e x( )

k x( ) Sr x( )

Qrf( ) Kf( )

(Increment, or iteration correction)Unbalanced forces (structure level) Increment of structure

displacements

Increment of elementdisplacements

Increment of sectiondeformations

Section restoringforces

Element restoringforces

Structure restoringforces

Deformation shapefunctions

Section model

Integration using thedeformation shape functions

Elem

ent S

tate

Det

erm

inat

ion

1

2

3

4

B x( )

B x( )

Page 67: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 43

shapes resulting from the stiffness decrease in the element non-linear zones. Note that, “cor-

rect” solution shall be understood within the assumptions of the section formulation and the

behaviour law.

2.4.3.2 Flexibility method

In the flexibility method, no displacement shape functions are assumed and the formulation is

cast in the element space without rigid body modes. As stated before, the element equilibrium

conditions (for the case of no element applied loads) lead to the total and incremental force

fields expressed by

(2.22)

where contains the force interpolation functions given by Eqs. (2.15).

The linearized section constitutive law can be also written in the inverse form of Eq. (2.18) as

(2.23)

Note that, as well as the operator, also the local flexibility and stiffness matrices

depend only on the section formulation. Since the Timoshenko beam formulation has been

adopted for the element developed in the present work, the corresponding matrix is given

in Appendix A, for the particular case of linear elastic behaviour. The elastic section flexibility

matrix consists only of diagonal constant terms; if non-linearity is to be included, then non-

constant terms do appear corresponding to the non-linear deformation components.

The compatibility between end displacement increments and the local section deforma-

tions is conveyed by application of the virtual force principle

(2.24)

upon an initial state of compatible deformations. Substituting Eqs. (2.22) and (2.23) in Eq.

(2.24), which must hold for an arbitrary choice of the virtual force vector , the flexibility

relationship at the element level is derived , where the flexibility matrix is

given by

S x( ) b x( ) Q⋅= and ∆S x( ) b x( ) ∆Q⋅=

b x( )

∆e x( ) f x( ) ∆S x( )⋅=

∂ f x( ) k x( )

f x( )

∆u

∆e x( )

δQT ∆u⋅ δST x( ) ∆e x( )⋅ xd0

L

∫=

δQ

∆u F ∆Q⋅=( )

Page 68: Seismic Assessment of Reinforced Concrete Frame Structures

44 Chapter 2

(2.25)

Note that Eq. (2.24) can be readily modified to obtain the element displacements in terms of

section deformations as given by Eq. (2.16).

The stiffness matrix is obtained by inversion and is then transformed to the flexible

element with rigid body modes by

(2.26)

In turn, the stiffness matrix in the element local space is then obtained by

(2.27)

upon which another reference system transformation must be performed to obtain the matrix

ready to be assembled in the global structural stiffness matrix .

This method elicits the following comments:

• the equilibrium at the element and section levels is strictly ensured by Eqs. (2.22), when no

element loads are included, because these equations remain unchanged regardless of the lin-

ear or non-linear behaviour; this means that the force interpolation functions are

always exact, contrarily to the deformation shape functions , which are only strictly

valid for linear behaviour of prismatic members;

• however, the above referred advantage of the flexibility method is counteracted by signifi-

cant difficulties on the numerical implementation, which have surely contributed to the

strong preference by displacement based methods in numerical modelling so far; actually,

finite element analysis programs are typically designed for imposing displacements (rather

than forces) at the member ends and for evaluating the corresponding resisting forces

(instead of displacements), for which deformation shape functions are usually required.

This final comment highlights the fact that the flexibility element state determination cannot be

directly included in the sequence shown in Figure 2.7. Actually, it requires more steps than the

stiffness case and an iterative scheme has to be enforced at the element level, as described in

the next section.

F bT x( ) f x( ) b x( )⋅ ⋅ xd0

L

∫=

u

e x( )

K F 1–=( )

Kf Tb K TbT

⋅ ⋅=

Ke

Ke Tr Kf TrT

⋅ ⋅=

KE KG

b x( )

B x( )

Page 69: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 45

2.4.4 The element state determination

2.4.4.1 General comments

The flexibility element state determination scheme is extensively described by Spacone et al.

(1992) and Taucer et al. (1991), in the flexibility method context and also in the framework of

the two field mixed method (Zienkiewicz and Taylor (1989)) as a more general formulation.

Both methods lead to the same result concerning the element force-displacement relation,

although, according to the referred authors, the latter method leads to a more consistent imple-

mentation of the most difficult task, i.e., the element state determination.

In the present work the intrinsic formalism of the mixed method was deemed unnecessary,

because, in our opinion, a good understanding of that formulation requires a previous knowl-

edge of the state determination sequence. After all, that is for sure the reason why the presenta-

tion of the mixed method approach in Spacone et al. (1992) is preceded by the description of

the state determination.

The element state determination in classical stiffness based computer programs consists in the

evaluation of the element stiffness matrix and restoring force vector . In the flexibil-

ity method the stiffness matrix can be readily obtained from the flexibility one as described

above, but the force vector , corresponding to given element displacements , cannot

be directly obtained due to the lack of deformation shape functions; therefore, a special proce-

dure has to be adopted, which is better understood in the context of the non-linear analysis

scheme.

The non-linear analysis algorithm at the structure level often consists of the classical incremen-

tal-iterative process schematically shown in Figure 2.8-a). The outer process refers to the

application of external loads, performed in a sequence of increments, and is denoted by super-

script k. The inner process, identified by superscript n, consists on the iterative process (of

Newton-Raphson type) necessary for completion of each load increment k.

The specific features of the present formulation, for which no explicit displacement shape

functions are available, require another iterative process at the element level for the state deter-

mination within each Newton-Raphson (N-R) iteration n. This innermost process, labelled

with superscript j and also included in Figure 2.8, yields the element restoring forces corre-

Kf( ) Qrf( )

Qrf( ) uf( )

Qn

Page 70: Seismic Assessment of Reinforced Concrete Frame Structures

46 Chapter 2

sponding to the element displacements . Figure 2.8-b) shows the internal iteration evolution

in terms of element forces and displacements for each N-R iteration n, while Figure 2.8-c)

schematically illustrates the corresponding development in terms of section forces and defor-

mations. In the present work, only this innermost iterative process (internal iteration scheme)

will be described with some detail in terms of the reduced variable space , which can be

related with the complete space ones by Eqs. (2.11) and (2.13).

Figure 2.8 State determination of flexibility based elements: from the global to the local level

un

Q u( , )

Qf uf( , )

Q

A

B

D

EFC

A

B

DC

A

B

D

EFC

S

Element loop ( j )for n-th N-Riteration

Internal loops ( j ) for each N-R iteration

ELEMENT

SECTION

un-1 uun un+1 un+2 uk

en-1 en en+1 e

Qn

Sn

A

B

D

E

F(load increment k)

A, B, D constitute the Newton-Raphson iteration n

STRUCTURE

uG( )n

uG( )n 1+

QG( )k

QG

uG( )n 1–

a)

c)

b)

uG

QG( )n

QG( )n 1–

QG( )k 1–

uG( )k

ej

Sj

Sn-1

Qn-1

Page 71: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 47

This special procedure in the flexibility formulation aims at satisfying both force equilibrium

and deformation compatibility with the imposed displacements at the element level. Therefore,

the procedure has to account for the possible existence of element applied loads, since it modi-

fies both the equilibrium conditions and the distribution of deformations along the element.

Thus, for the sake of simplicity, two situations are distinguished concerning applied loads: the

case of nodal forces only, which is introduced first in 2.4.4.2, and the case with element loads

also included, which is explained in 2.4.4.3 as an extension of the previous one.

2.4.4.2 Nodal forces

The special procedure for element state determination is better understood by means of a pre-

liminary comparison with the usual process of classical stiffness based state determination.

Referring to Figure 2.7, the non-existence of explicitly defined displacement shape functions

for flexibility based elements does not allow to perform tasks 2 and 4. Instead, the transition to

section level must be done in terms of forces, following the sequence described in Figure 2.9,

where the first and the last tasks are labelled with 1 and 4, respectively, in order to highlight

the correspondence with the scheme of Figure 2.7.

Figure 2.9 Flowchart for the element state determination of flexibility based elements

The increment of end section forces (in brief element forces) is first computed from the incre-

ment of element displacements using the last stiffness matrix; then, with the force shape func-

tions, the increment of section forces can be obtained. At this point, section deformations

should be calculated, but difficulties arise because non-linear section force-deformation rela-

Incr. of element forces

Increment or residuals of element displacements

Incr. of section deformations

Section restoring forces,flexibility and residual forces

Section residual deformations

Force shapefunctions

Section model

Integration weighted

Incr. of section forcesCurrent section flexibility

Updated section flexibility

by force shape functions

Current element stiffness

Element restoring forces4

1

Page 72: Seismic Assessment of Reinforced Concrete Frame Structures

48 Chapter 2

tionships are usually expressed in terms of deformations. Additionally, since the section

applied forces modify the section stiffness, the associated element stiffness matrix is no more

the one used to obtain these forces. The problem is solved by recourse to a non-linear iterative

procedure at the element level, which is stopped when element residual displacements vanish,

as required by nodal kinematic compatibility.

The process evolution, at both element and section levels, is illustrated in Figure 2.10 for the

generic n-th Newton-Raphson iteration where the element resisting forces , in correspond-

ence with the element displacements , are to be obtained. Figure 2.10 also

includes the initial conditions (j=1) and the main expressions involved; for section variables

and expressions, the reference to the generic abscissa (x) was suppressed for simplicity.

The process starts at point A with the application of the first element force increment

, obtained from the given element displacement increments and using the

initial tangent flexibility matrix in the inverse form; the element state point moves to position

B. The section force increments are computed using the force shape functions

and the prediction of section deformation increments is given by ,

using the local flexibility matrix compatible with the previously used element flexibility

matrix. Both force and deformation vectors can be updated to and , respectively,

and the section state is transferred to point B.

The section model allows for the restoring forces and the new tangent flexibility

to be obtained. However, force residuals are likely to appear until the state point B does not

reach the model curve. Since the equilibrium has to be satisfied, these force residuals

cannot be allowed and so the section has to deform more; using the section

force-deformation relation linearized about the updated state, the residual section deformations

can be given by .

If these residual deformations were allowed to take place, then, also at the element level, resid-

ual displacements would appear and be computed by , according to

Eq. (2.16). In such case, at both the section and the element levels, the point B (corresponding

to the final state for iteration j=1) would move to position B’; however, the kinematic compat-

ibility at the element level prevents displacements to go beyond .

Qn

un un-1 ∆un+=

∆Q1 F0[ ]1–

∆u1⋅=

∆S1 x( ) b x( )

∆e1 x( ) f0 x( ) ∆S1 x( )⋅=

S1 x( ) e1 x( )

Sr1 x( ) f1 x( )

S1 x( ) Sr1 x( )–( )

∆er1 x( ) f1 x( ) S1 x( ) Sr

1 x( )–( )⋅=

∆ur1 bT x( ) ∆er

1 x( )⋅ xd0

L∫=

un

Page 73: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 49

Figure 2.10 Details of element and section state determination for flexibility based elements

This means that corrective element forces must be applied in order to restore compatibility by

pushing the point B’ back to the allowed displacements. Based on the updated element tangent

flexibility matrix (see Eq. (2.25)), corrective forces can be obtained by ,

constituting the increment of element forces for the second iteration (j=2); thereby, the

state point moves now to position C. Again, through the force shape functions, the section

force increments are obtained and applied starting from the previous

force level ; the corresponding deformation increments are superimposed

Q

u

A

B

D

C

A

B

DC

e

S

ELEMENT

SECTION ...(x)

B’

C’

B’C’

Sr1 Sn

Sn-1

∆S1

∆S2

∆S3

en-1 en∆e1

f0

f1

f2

un∆un

QnQn-1

∆Q1

∆Q2

∆Q3

F0F1

F2 Initial conditions ( j=1 ): F0 Fn-1= ∆u1 ∆un=

Generic iteration:

∆Qj Fj-1[ ]1–

∆uj ∆urj-1–( )⋅=

Qn Qn-1 ∆Qj

j 1=

converg.

∑+=

Initial conditions ( j=1 ): f0 fn-1= ∆er

0 0=

Generic iteration:∆Sj b ∆Qj⋅=

∆ej fj-1 ∆Sj⋅ ∆erj-1+=

Sn Sn-1 ∆Sj

j 1=

converg.

∑+=

en en-1 ∆ej

j 1=

converg.

∑+=

S1

S2

S3

e1 ...∆er

j fj Sj Srj–( )⋅=

∆urj-1 bT x( ) ∆er

j-1⋅ xd0

L

∫=

∆uj 1> 0=

un-1

∆er2

∆ur1

∆ur2

∆er1

F1 F1[ ]1–

∆ur1–( )⋅

∆Q2

∆S2 x( ) b x( ) ∆Q2⋅=

S1 x( ) f1 x( ) ∆S2 x( )⋅

Page 74: Seismic Assessment of Reinforced Concrete Frame Structures

50 Chapter 2

to the previous residual deformations , thus moving the section state point to position

C. At this stage the new restoring forces and tangent flexibility matrices are com-

puted and, the process is repeated as for the first iteration: residual deformations are calculated

and then integrated to yield new residual displacements, which in turn generate new element

corrective forces to start the third iteration and so on. The process stops when an adopted

measure of residuals (in terms of section forces or deformations, or element displacements for

example) is less than a pre-defined tolerance.

In Figure 2.10, convergence is supposed to be achieved at point D, which gives the element

restoring forces for the given displacements ; from then on, the Newton-Raphson proc-

ess, at the structure level, can proceed to the next iteration n+1.

Two aspects of this process should be underlined:

• at the element level, after the given initial displacement increments have been applied (j=1),

the displacements remain fixed while forces are repeatedly corrected; for this reason the

points B, C and D, representing the element state during iteration loops j>1, lie in the same

vertical line as shown at the top of Figure 2.10;

• at the section level, both deformations and forces are continuously updated and so, the state

points B, C and D cannot lie in the same vertical line; this complies with the force equilib-

rium that must be strictly satisfied in all sections, through the exact force shape functions,

and with the kinematic compatibility conditions to be satisfied in integral form as expressed

by Eq. (2.16).

2.4.4.3 Element loads

Element loads (concentrated or distributed along the span) cannot be included in the same way

as in the usual stiffness based methods, since the displacement shape functions (from which

the equivalent nodal forces are usually obtained) are not explicitly known and do not remain

constant throughout the non-linear process.

Therefore, for each load increment a set of equivalent nodal forces due to element loads is

implicitly computed and included at the element level by an iterative process (similar to the

one used when only nodal loads are present), which ensures that both the equilibrium and the

compatibility conditions are satisfied.

∆er1 x( )

Sr2 x( ) f2 x( )

Qn un

Page 75: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 51

As in Figure 2.10, the process sequence at both the element and the section levels is shown in

Figure 2.11, for the generic n-th N-R iteration of the k-th load increment, and follows almost

the same steps as proposed by Taucer et al. (1991). It is assumed that a prescribed increment

of element loads is included simultaneously with other nodal loads and remains constant

during the N-R scheme.

Figure 2.11 Details of element and section state determination for flexibility based elements

with the application of element loads.

When each step starts (n=1), the whole structure, considered as a discrete system in the global

nodal variable space, does not “see” the element loads to be applied in that step. Thus, the

process begins with the application of the nodal load increments only, generating increments of

nodal displacements which are then transferred to the element level as described before.

∆p

Q

u

A

B

D

C

A

B

DC

e

S

B’’C’

B’C’

Sr1 Sn

Sn-1

∆SQ1

∆S2

∆S3

en-1 en

∆e1

∆er2

f0 f1

f2

un∆unun-1

QnQn-1

∆Q1

∆ur1

∆ur2

∆Q2

∆Q3

F0

F1F2

S1

S2

S3

e1 ...

B’’∆Sp

∆S1

B’

∆ur1( )Q ∆ur( )p

∆er1( )Q

∆er1

∆er( )p

ELEMENT

SECTION

Page 76: Seismic Assessment of Reinforced Concrete Frame Structures

52 Chapter 2

Exactly as in Figure 2.10, the element and section state points start from position A up to posi-

tion B (see Figure 2.11), where the section state determination takes place and the local equi-

librium has to be controlled. Here, the main difference appears with respect to the process

described before: according to Eq. (2.14), part of the section forces are due to the element

applied loads (denoted by in Figure 2.11) which have to be superimposed to those aris-

ing from nodal forces (referred to as , where the superscript stands for the internal iter-

ation j=1 and the subscript identifies these internal forces source). Hence, with the total

increment of internal force for the first iteration , the section

force is updated to the level , not coincident with that of point B.

When equilibrium is controlled at each section, it is apparent that unbalanced forces, given by

, include now an explicit term due the element loads. For the equilib-

rium to be satisfied, residual deformations are considered as before

, using the updated section flexibility ; residual deformations

consist of a term due to nodal loads and another due to element loads , as

shown in Figure 2.11. These residuals enforce the section point B to move to the new position

B” at a different level from the point B’ associated only with the nodal load contribution.

At the element level, displacement residuals can be obtained in the same way as in 2.4.4.2,

which allows for the computation of corrective forces, using the updated flexibility matrix ,

in order to start the second iteration. The application of these element loads moves the

state point B” to the position C, restoring displacement compatibility. Again, at the section

level, the corrective increment of forces is applied from the previous

force level, which, in this case, corresponds to point B”. From this stage on, the process fol-

lows exactly the same steps as for the case with only nodal loads, since the element loads are

already applied and only corrective nodal forces remain to be imposed until the displacement

compatibility is reached.

At the convergence stage (at points D in Figure 2.11) the section forces, deformations and flex-

ibility are obtained, taking into account the whole increment of element loads . Also at the

element level, the restoring forces are such that the imposed displacements are satisfied while

accounting for the influence of . This means that for the next Newton-Raphson iterations

(n>1) the increment of element loads must not be included again, since the element state deter-

mination starts from a stage in which all the increment is already considered.

∆Sp x( )

∆SQ1 x( )

∆S1 x( ) ∆SQ1 x( ) ∆Sp x( )+=( )

S1 x( )

∆Su1 x( ) S1 x( ) Sr

1 x( )–=

∆er1 x( ) f1 x( ) ∆Su

1 x( )⋅=( ) f1 x( )

∆er1 x( )Q( ) ∆er x( )p( )

F1

∆Q2

∆S2 x( ) b x( ) ∆Q2⋅=

∆p

∆p

∆p

Page 77: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 53

It should be emphasized that, the influence of element applied loads in the residual displace-

ments is taken into account only after the state determination is performed. Actually, one could

think of performing the state determination for a predictor deformation increment

including the contribution of ; obviously, both the resisting force and the flexi-

bility would be different, as well as the section residuals . However, due to the

influence of , the integration of deformation residuals would not lead to the adequate

displacement residuals, since they refer to a deformation state which is no longer coherent with

the target displacements . Therefore, if such a strategy were adopted, a further correction

would be needed at the residual displacement level, in order to refer them to the right target

displacements. In our opinion, this leads to a cumbersome process for residuals evaluation, and

thus, the option of including only after the state determination appears to be prefera-

ble.

The scheme for this sequence of incremental element and nodal load application is illustrated

in Figure 2.12 for the common case of planar frame structures where only vertical uniformly

distributed loads are considered together with other nodal forces acting in the frame plane.

The load increment k consists of global nodal forces or displacements and of ele-

ment loads as indicated. For simplicity the initial state of the increment is associated with

the superscript 0, although it actually corresponds to the final state of the previous step k-1. At

the element level, the reduced variable space includes only end section bending moments

and rotations .

For the first N-R iteration (n=1) the whole increment is applied simultaneously with the

rotation increments (obtained directly from ), starting from a stage with

element loads and end section moments and rotations. It is assumed that does not

vary with displacement corrections.

During the first internal iteration (j=1), the increments are applied, through

the corresponding first approximation of moments , and the increment is included

in the residual corrective phase, as explained above. For the next internal iterations (j>1) the

level of element loads is already updated to and only corrective end section

moments are due. The same happens for the following N-R iterations (n>1).

∆e1 x( )

∆Sp x( ) Sr1 x( )

f1 x( ) ∆er1 x( )

∆Sp x( )

un

∆Sp x( )

∆QG ∆uG( , )

∆p

M

θ

∆p

∆θn=1 ∆uG( )n=1 p0

M θ( , )0 ∆p

∆θj=1 ∆θn=1=

∆Mj=1 ∆p

p p0 ∆p+=

∆Mj

Page 78: Seismic Assessment of Reinforced Concrete Frame Structures

54 Chapter 2

Figure 2.12 Sequence for the application of incremental element and nodal loads

Given the fact that substantial differences appear in the process only for the first N-R and inter-

nal iterations (n=1 and j=1), the detailed evolution for that stage is shown in Figure 2.13 for

arbitrary distributions of moment, flexibility and curvature.

∆θn

1 2

p0

QG uG( , )0

1 2

∆p

∆QG ∆uG( , )

1 2

p

QG uG( , )

+ =

Load increment k ;

NEWTON-RAPHSON ITERATIONS

p0

∆p

+

+

M θ( , )0

∆θn=1

p0

∆p

+

+

M θ( , )0

∆Mj=1

p p0 ∆p+=

+M θ( , )n-1

INTERNAL ITERATIONS

+M θ( , )j-1

∆Mj

p p0 ∆p+=

+M θ( , )j-1

∆Mj

p p0 ∆p+=

j 1≥j 1>j 1=

n 1>n 1=

Initial state 0 k 1–( )≡

Page 79: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 55

The distinct tasks are identified, such as the predicting and updating phase, the section state

determination, the equilibrium checking and the local residuals computation. Only flexural

non-linearity is considered and, for simplicity, starting only from the left end section. A clear

distinction is made between internal force distributions due to element loads and nodal loads,

and it is evidenced that effects of are included only when section equilibrium is checked.

Figure 2.13 Details of the element state determination for first internal iteration of the first N-R

iteration, in the presence of element loads

∆p

+

+ =

ELEMENT LEVEL

+

+ =

+ =

=

-MQ

0 x( )

M0 x( )

Mp0 x( )

f0 x( )

∆MQ1 x( )

∆ϕQ1 x( )

MQ1 x( )

ϕ1 x( )

MQ1 x( )

∆Mp x( )

ϕ0 x( )

M1 x( )

Mr1 x( )

f1 x( )

∆Mu1 x( )

∆ϕr1 x( ) f1 x( )∆Mu

1 x( )=

∆θr1 x( )

Sect

ion

Stat

eDe

term

inat

ion

Section Equilibrium

j 1=

n 1=N-R iteration

Internal iteration

∆ϕQ1 x( ) f0 x( )∆MQ

1 x( )=

F1

Section Residuals

Predicting and updating phase

Load increment k ; Initial state 0 k 1–( )≡

Page 80: Seismic Assessment of Reinforced Concrete Frame Structures

56 Chapter 2

Figure 2.14 shows the same type of evolution, but for the subsequent internal iterations (j>1),

in order to give a comprehensive view of how the process develops. The same scheme is also

valid for all the successive internal iterations of the (n>1) N-R iterations; it should be noted

that no more effects of are considered at this stage (n>1) and that, during the predicting

phase, the previous residual deformations are included.

Figure 2.14 Details of the element state determination for the (j>1) internal iterations of the first

N-R iteration, in the presence of element loads

∆p

ELEMENT LEVEL

+ =

+ =

=

+

M1 x( )

f1 x( )

∆MQ2 x( )

∆ϕQ2 x( )

M2 x( )

ϕ2 x( )ϕ1 x( )

M2 x( ) Mr2 x( )

f2 x( )

∆Mu2 x( )

∆θr2 x( )

Sect

ion

Stat

eDe

term

inat

ion

Section Equilibrium

∆ϕr1 x( )

N-R iteration

Internal iteration j 1>

∆ϕQ2 x( ) f1 x( )∆MQ

2 x( )=

∆ϕr2 x( ) f1 x( )∆Mu

1 x( )=

F2

Section Residuals

Predicting and updating phase

n 1=

Load increment k ; Initial state 0 k 1–( )≡

Page 81: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 57

At each N-R iteration, after the element restoring forces are obtained, they have to be

transformed into the flexible element local space of variables. Eq. (2.11) expresses the required

transformation, where the element loads have now a relevant contribution to the force

components not belonging to the element reduced space. Additionally, these element loads

shall be in accordance with those used in the state determination for the N-R iteration. There-

fore, if total restoring forces are to be transferred, then also the total load shall be used;

on the contrary, in the case of restoring force increments , then also the increment

must be considered in the transformation.

The restoring force vector is transferred to the element global space through the rela-

tions given in 2.4.2, and assembled in the global nodal restoring force vector. Finally global

unbalanced forces can be obtained and, then, the next iteration starts.

It is worth mentioning that vectors contain the nodal forces equivalent to the applied ele-

ment loads for the present state of the element. They have been implicitly obtained during the

internal iterative process and, obviously, they change when the flexibility distribution changes.

The special case, in which element loads are applied only in the first increment without any

other global nodal forces, helps to understand how the process develops. In fact, these condi-

tions mean that, for the first N-R iteration (n=1), the element state determination is performed

aiming at null element displacements in order to satisfy nodal kinematic compatibility. How-

ever, nodal element restoring forces are not null and are likely to generate non-zero global

nodal forces. Since these must be zero, unbalanced forces do appear and generate displacement

corrections for the next N-R iterations (n>1). When convergence is reached, the final state of

the structure is characterized by zero nodal loads associated with non-zero nodal displacements

due to the element loads.

A slightly different version for the application of element loads was proposed by Spacone

(Spacone (1994) and Spacone et al. (1996)) and is briefly presented next because of its clarity.

For a given N-R iteration n, should the displacement increments be applied simultane-

ously with an element load increment , the element force increments for the first internal

iteration (j=1) must be given by

(2.28)

Qn

Qpf( )

Qn p

∆Qn ∆p

Qf( )n

Qf( )n

∆un

∆p

∆Qj=1 Kj=0 ∆un⋅ ∆Qp+=

Page 82: Seismic Assessment of Reinforced Concrete Frame Structures

58 Chapter 2

where and is the vector of element forces necessary to ensure the

desired displacement increment . Actually, represents the element fixed end forces

due to , conveniently updated for the current state of the element.

According to Eq. (2.14), the section force and deformation increments are given by

(2.29)

(2.30)

The application of the virtual force principle, leads to the following expression

(2.31)

whose validity requires that

(2.32)

Therefore, can be computed before the element iterations start and be included in the sec-

tion incremental forces at the first iteration (j=1) according to

(2.33)

The section state determination is carried out for the predictor deformations

(2.34)

whose compatibility with the target displacements is ensured by the way was

obtained. For the subsequent iterations (j>1) the algorithm proceeds as before, i.e., only cor-

rections due to nodal forces have to be made until displacement compatibility is reached.

This procedure appears to treat more clearly the element load application, because it highlights

more rationally the fixed-end forces and avoids the somewhat tricky process of considering

those load effects only after the section state determination.

Kj=0 Fj=0[ ]1–

= ∆Qp

∆un ∆Qp

∆p

∆S x( )j=1 b x( ) ∆Qj=1⋅ ∆Sp x( )+=

∆e x( )j=1 f x( )j=0 ∆S x( )j=1⋅=

∆un bT x( ) f x( )j=0 b x( ) Kj=0 ∆un⋅ ∆Qp+[ ]⋅ ∆Sp x( )+{ }⋅ ⋅ xd0

L

∫=

∆Qp K– j=0 bT x( ) f x( )j=0 ∆Sp x( )⋅ ⋅ xd0

L

∫⋅=

∆Qp

∆S x( )j=1 b x( ) Kj=0 ∆un⋅ ∆Qp+[ ]⋅ ∆Sp x( )+=

e x( )j=1 e x( )j=0 f x( )j=0 ∆S x( )j=1⋅+=

un ∆Qp

Page 83: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 59

2.4.4.4 Remarks on the global non-linear algorithm

The element state determination described so far has been cast within a global N-R algorithm

whereby the element incremental forces (as well as the incremental section forces and

deformations) are calculated from iterative displacements rather than incremental ones. In

other words, iteration n is performed by considering the iterative correction starting from

displacements and forces at the end of the previous iteration n-1, for which equilibrium has not

been reached yet.

For materials having path-dependent properties, such as crack opening and closing and crush-

ing in concrete, it is emphasized (Argyris et al. (1978)) that, along with the adoption of small

increments, the evaluation of incremental restoring forces (or stresses) be done from incremen-

tal deformations (i.e., those referred to the step beginning as an equilibrated state). This

requires another scheme as shown in Figure 2.15 (slightly different from that of Figure 2.8)

where the state determination is performed for a total displacement increment referring to

the displacement at the step beginning, which is transformed into element forces

with the initial stiffness matrix of that step. Section forces and deformations are then succes-

sively obtained and corrected following the same procedure as before.

The whole process for the element state determination remains almost unchanged, provided

the updating of force and deformation vectors (at both element and section levels) is performed

only when convergence is reached also at the structure level.

However, a specific remark is due concerning the application of element loads. Actually, for a

given step where an increment is to be applied, the fact that every N-R iteration n is

referred to the step beginning, requires the contribution of to be included in the first inter-

nal iteration j for all N-R iterations. This arises from the fact that this algorithm does not

“memorize” the loading history from preceding iterations, but only from the previous step

increments. Therefore, in the element load application sequence schematically illustrated in

Figure 2.12, the leftmost part referring to the first N-R iteration becomes valid for any

nth N-R iteration, whereas the rightmost part is not applicable. Accordingly, the illustrative

details of the element state determination included in Figures 2.13 and 2.14 are also valid for

any nth N-R iteration rather than only for the first one.

∆Qn

∆un

∆un

∆u0n

u0 uk-1=( )

∆p

∆p

n 1=( )

Page 84: Seismic Assessment of Reinforced Concrete Frame Structures

60 Chapter 2

Figure 2.15 State determination of flexibility based elements for displacements corrections rel-

ative to the step beginning

Q

A

B

DC

S

Element loop ( j )for n-th N-Riteration

Internal loops ( j ) for each N-R iteration

ELEMENT

SECTION

u0 uu1 un uk

e0 en e

Qk

Sn

uG( )2… uG( )

n…

QG( )k

QG

uG( )1

c)

b)

uG

QG( )n

QG( )1

QG( )k 1–

uG( )k

e1

S1

S0 Sk-1=

QG( )2

uG( )k 1–

uG( )0

=

(load increment k)

STRUCTUREa)

Q1

Qn

Q0 Qk-1=

Bn

B1Cn

C1 Dn

D1

Dn

D1

D2

B

A

A

Bn

B1Cn

C1DnD1

A

Page 85: Seismic Assessment of Reinforced Concrete Frame Structures

MODELLING OF REINFORCED CONCRETE FRAME STRUCTURES - STATE OF THE ART 61

2.4.4.5 Control sections and numerical integration

All the element integrations, for calculation of both the flexibility matrices and the residual

displacement vectors, are numerically performed. A set of control sections is pre-defined to be

monitored during the analysis, whose location is chosen according to the expected distribution

of non-linearity along the element. Since the inelastic behaviour often concentrates at member

ends, control sections shall also concentrate in these zones.

Additionally, the end sections have also to be monitored and, therefore, the Gauss-Lobatto

scheme appears more adequate (Spacone et al. (1992)) than the Gauss-Legendre method. The

number of control sections to be considered is chosen, as usual, according to the degree of pol-

ynomials to be integrated; it is recalled that, the Gauss-Lobato scheme with m integration

points leads to exact integration of polynomials of degree up to (2m-3).

2.5 Concluding summary

The present chapter has described the framework for the hysteretic behaviour modelling of

reinforced concrete frame structures, starting with a preliminary identification of the main fea-

tures of member behaviour under monotonic and cyclic conditions. Particularly, cyclic load

effects such as the stiffness degradation (for unloading and reloading stages), the pinching

effect and the strength deterioration have been presented in order to highlight the most relevant

aspects based on which the phenomenological models are derived.

A good compromise between details of structural response and model manageability can be

obtained with member models, whose evolution over the last three decades has been presented

in a historical perspective. The major requirements for member modelling have been identi-

fied, viz the hysteretic model and the element model.

The first developments of member models in uniaxial bending conditions have been recalled,

namely with reference to the two-component and the one-component (or point hinge) models,

and some of the most relevant phenomenological hysteretic models have been briefly referred.

Member models with distributed inelasticity gained increasing acceptance for non-linear frame

analysis upon recognition of the point-hinge model limitations to adapt to actual force and

stiffness distributions along the member. Thus, the evolution has been described, from the ear-

Page 86: Seismic Assessment of Reinforced Concrete Frame Structures

62 Chapter 2

liest attempt by Otani (1974) to the recent proposal in the flexibility context by Taucer et al.

(1991) and Spacone et al. (1992), reflecting the continuous effort of researchers to develop

member models that closely follow the spread of inelastic effects along the member while

accounting for the various sources of non-linear phenomena.

The advent of flexibility based formulations has been highlighted and emphasis has been put

on its adequacy for frame member modelling, as a rather suitable formulation to account for

inelasticity spread, to accommodate different types of section model and to easily incorporate

the association in series of non-linear sub-elements.

Since the member model developed in the present work belongs to the family of flexibility

based models, attention has been focused on the general flexibility formulation for beam-col-

umn elements as described in previous work. The main issues of the theoretical bases have

been addressed and the steps for the element state determination in flexibility models (thus in

the absence of deformation shape functions) have been presented, first for the case of nodal

loads only and then for the element applied loads also; additionally, its insertion in classical

stiffness based finite element algorithms has been discussed.

Page 87: Seismic Assessment of Reinforced Concrete Frame Structures

Chapter 3

FLEXIBILITY BASED ELEMENT WITH

MULTI-LINEAR GLOBAL SECTION MODEL

3.1 General comments and innovative features

The importance of adequately modelling non-linear behaviour has been discussed in Chapter

2, where member models have been recognized particularly suitable for seismic non-linear

analysis of reinforced concrete frame structures.

The state-of-the-art review presented in the previous chapter illustrated the great amount of

work developed in frame modelling, from which very powerful models became available for

non-linear behaviour simulation. The identification of the involved phenomena and the succes-

sive proposals to take them into account, led to models with remarkable level of sophistication.

Despite their unquestionable value, highly sophisticated models may easily become incompat-

ible with the needs of massive calculations often encountered in seismic analysis.

In this context, a major concern of the present work has been the research and development of

a modelling tool with a moderate level of sophistication, particularly in what concerns non-lin-

ear behaviour. Thus, while retaining the advanced features of recent sophisticated member

models (Taucer et al. (1991), Spacone et al. (1992), Spacone (1994)), a new element model has

been sought, which provides robustness, efficiency/economy (in terms of discretization and

computation time) and the ability to trace the non-linear behaviour right from the early stages

(such as the post-cracking/pre-yielding phase, typically corresponding to serviceability limit

states) up to advanced inelasticity development due to large ductility demands. Particularly, the

non-linear behaviour in the pre-yielding phase (very often neglected due to its reduced influ-

ence in extensive yielding stages) has been recently reported as important for an adequate

Page 88: Seismic Assessment of Reinforced Concrete Frame Structures

64 Chapter 3

assessment of the structural dynamic characteristics (Fardis and Panagiotakos (1997), Calvi

and Pinto (1996)) and of the seismic response for serviceability states (as required in modern

seismic design codes, e.g. EC8).

Therefore, in order to contribute for member model development or improvement, the follow-

ing basic issues were deemed necessary:

• The ability to adequately describe the member stiffness, in particular its evolution through-

out the various behaviour stages during the response; this is indeed an important issue in the

seismic analysis context, where a good assessment of structural frequencies of vibration has

a direct influence on the dynamic response.

• The non-linear behaviour modelling should not go beyond the global section level, so as to

ensure a reasonably low cost of seismic analysis computations.

• In order to achieve discretization and computation economy and ease of usage (preferably

comparable to that of linear analysis with classical beam elements), each structural member

(beam, column or inter-storey piece of wall) should be modelled, as much as possible, by

only one element.

• The number of control sections in the element should be reduced to the minimum required,

yet ensuring an adequate description of properties along the member and a sufficiently

accurate simulation of the response.

• The possibility of incorporating new developments, allowing for a better adjustment to the

actual member behaviour, namely by including other non-linearity sources not yet consid-

ered (e.g. effects of bar slippage, inelastic shear).

• The ease of incorporation, without major pre-requirements, in classical non-linear analysis

algorithms typically based on standard or modified Newton-Raphson methods.

The general flexibility element introduced in 2.4, is an excellent means for closely tracing out

the structural non-linear response and, with the adoption of the fibre model (Taucer et al.

(1991)) for reproducing the local section behaviour, a very powerful modelling tool is

obtained. However, such a methodology can hardly fulfil the requirement of efficiency and

economy for the analysis of building structures with large number of elements, specially when

only the global element behaviour is of interest.

The use of global models to describe section behaviour appears as a logic and natural option in

order to improve computation time performance of the flexibility element formulation, and, as

Page 89: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 65

mentioned above, it has been adopted in a previous work (Spacone et al. (1992)) where a dif-

ferential section constitutive law is considered. However, in that proposal, the behaviour law

still had to be controlled at several control sections defined a priori and the effects of cracking

development were not properly taken into account because sections were assumed cracked

from the beginning.

To some extent, the need of several sections, chosen according to an expected element behav-

iour and distribution of non-linearity, is similar to the usual mesh refinement in classical stiff-

ness based finite elements. Therefore, in terms of time economy, the advantage of flexibility

elements over the stiffness ones is not evident, since control sections may be specified in zones

where non-linearity is not likely to appear.

In the present study, a global section constitutive law is also adopted; however, an effort is

made to restrict the number of pre-defined control sections, yet adequately tracing out the dam-

aging process along the element. For this purpose a multi-linear step wise model is adopted,

based on a trilinear envelope curve (for each direction of deformation) along with additional

hysteretic rules. The cracking and yielding section stages are taken into account and the effects

of other RC typical phenomena are also included, such as stiffness and strength degradation

and pinching (due to crack closing, bond slip or non-linear shear deformations). It will be

shown that such a behaviour modelling option, together with some simplifying assumptions,

allows for the definition of special control sections which move during the loading process and

reduce the need for pre-defined locations.

After the statement of some basic assumptions and the adjustment of convention and notation

in 3.2, the adopted global section model is presented in 3.3 referring to a moment-curvature

Takeda-type model based on a trilinear skeleton curve. Control sections and distinct zones con-

sidered along the element are introduced in 3.4; these are assumed of two types, viz the fixed

and the moving ones, the former consisting of the element end sections and a mid-span one,

while the latter stand for the yielding, cracking and null-moment sections. In view of the mov-

ing section development, the way of considering element applied loads is also discussed in 3.4,

where a simplifying assumption is introduced. The behaviour of control sections is addressed

in 3.5, where a modification is considered in the trilinear model, in order to make possible the

approximate control of cracked zones by means of a special transition from uncracked to fully-

cracked behaviour.

Page 90: Seismic Assessment of Reinforced Concrete Frame Structures

66 Chapter 3

The element state determination is thoroughly detailed in 3.6. The flexibility distribution along

the element is presented and due account is taken of the influence of moving cracking sections

on the element flexibility matrix. Although based on the flexibility formulation, the specific

features of this element, arising from the combination of fixed and moving control sections,

require particular care on the evaluation of displacement residuals. The evolution of non-linear

behaviour in the element end zones is carefully addressed and an event-to-event scheme is

adopted for adequate control of residuals. Convergence criteria are stated and convergence

problems related with the cracking transition led to the adoption of a line-search scheme per-

formed inside the element iterative process.

The non-linear algorithm is summarized in 3.7, where the main steps of the general non-linear

scheme for static and dynamic analysis are first recalled to highlight the element state determi-

nation stage; then, the sequence of steps for that element stage is described according to the

preceding sections. The main features of the developed member model are recalled in 3.8.

3.2 Basic assumptions and remarks on convention and notation

The basic notation and conventions to be used next are essentially the same as adopted in 2.4

for the general flexibility formulation. However, some adjustments are introduced below, aim-

ing at notation simplification and better adaptation to the specific features of the element.

Referring to Figure 2.5, non-linear behaviour is only assumed in one bending plane (xz or xy),

while all the remaining deformation components are considered linear. This assumption is

acceptable if the effects of interaction between internal forces are negligible as in the case of

orthogonal beams in space frame structures. However, it can become a major drawback of the

formulation if, for example, columns are to be analysed in bi-axial bending with axial force

interaction. Note, however, that such limitation is not exclusive of the present modelling strat-

egy, since it mainly arises from the option for a global section model.

The previous assumption allows to omit the subscripts referring to directions of element and

section forces, since the relevant aspects for non-linear behaviour only refer to one bending

direction (by default assumed to be the y direction). Therefore, from now on, bending

moments, curvatures and rotations will be simply denoted, respectively, by M, ϕ, and θ; for the

related shear force and distortion the notation simplifies to V and β.

Page 91: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 67

The moving control sections depend on the moment distribution along the member and, there-

fore, element applied loads have to be pre-defined. In the present work these loads consist of a

uniformly distributed force vector (denoted by p) and a concentrated force vector (P) at the

span section H defined in 3.4. However, for the next developments, the relevant loads reduce

to those acting in the plane where non-linear behaviour is considered (thus, by default in direc-

tion z) and are simply denoted by p and P, respectively, for the distributed and concentrated

forces. If additional concentrated forces or different intensities of p need to be considered

along the member, a more refined discretization must be adopted.

By contrast with the general methodology introduced in 2.4.4.3, element load application will

be considered only in the first loading step and without any other load type. Actually, this cor-

responds to the common loading situation, for which the vertical static forces, due to self-

weight and live loads, are installed prior to other lateral actions (e.g. wind or earthquake

induced loading). Significant simplification of the numerical implementation can be achieved

with this option, but it is worth mentioning that, from the formal and conceptual standpoints,

the inclusion of element loads after the first step is straightforward.

The behaviour control is to be performed at the section level, in terms of generalized stress and

strain; thus, the internal force convention must be clearly defined and the section axis system

must be adjusted accordingly. For non-linear bending in plane xz (usually vertical) as assumed

by default, the section bending moments are considered positive when producing tensile strain

at the bottom fibres; the moment-curvature relationships are defined accordingly.

For the local section reference system assumed in 2.4, the above requirement is not satisfied,

since positive moments produce tensile strain at the top fibres. Thus, a rotation of the axis sys-

tem is introduced as shown in Figure 3.1, inducing a change of sign on the y and z components.

Figure 3.1 Adjustment of local section axis system

E1

zy

Mzs

ns ns2

xs xs2ns1

xs1ys

zs

Mys

zs2

ys2

ys1

zs1 E2

x

Page 92: Seismic Assessment of Reinforced Concrete Frame Structures

68 Chapter 3

The matrix and the vector of Eq. (2.14) are redefined accordingly, leading to

(3.1)

which hold for the internal force expressions in the rotated axis system, given by

(3.2)

and associated with the conjugate deformations denoted by .

Figure 3.1 also includes the end section axis systems associated with the respective positive

normal unit vectors ( or ), as commonly used in bending analysis. This allows to define

end section internal forces satisfying the above convention and to transform the forces and dis-

placements of the element reduced space into the local end section axis system.

Denoting by and the transformed forces and displacements, where the subscript es

stands for end sections, the following relations apply

(3.3)

as well as their inverse forms, since the matrix is invertible due to the orthogonal character

of the transformation. refers only to the reduced space degrees of freedom and is obtained

from the total matrix of the axis system transformation, thus having the expression

(3.4)

such that .

b x( ) Sp x( )

bs x( )

1– 0 0 0 0 00 0 0 1 L⁄ 0 1 L⁄0 0 1– L⁄ 0 1– L⁄ 00 1– 0 0 0 00 0 1 x L⁄–( ) 0 x– L⁄ 00 0 0 1 x L⁄–( ) 0 x L⁄–

= and Spsx( )

x–

zL L⁄ y+

– yL L⁄ z+

x–

x– yL L⁄( ) y+

x– zL L⁄( ) z+⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

Ss x( )

Ss x( ) bs x( ) Q⋅ Spsx( )+=

es x( )

ns1ns2

Q u( , )

Qes ues

Q Tes Qes⋅= ues TesT u⋅=

Tes

Tes

Tes

1– 0 0 0 0 00 1– 0 0 0 00 0 1 0 0 00 0 0 1 0 00 0 0 0 1– 00 0 0 0 0 1–

=

TesT Tes

-1 Tes= =

Page 93: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 69

Introducing Eqs. (3.3) in Eq. (3.2), the internal forces can be expressed in terms of

by

(3.5)

where . Therefore, the shape function matrix , relating the generic

section internal forces with those of the end sections in the rotated reference frames shown in

Figure 3.1, is given by

(3.6)

The expression equivalent to Eq. (2.16), in the present axis systems, is written

(3.7)

The advantage of working with these rotated reference frames is related to the fact that all

forces and deformations become directly compliant with the global section behaviour law.

However, transformations to the element axis system (x,y,z) have to be performed, since all the

remaining relations have been defined referring to it. Force and deformation transformations

are carried out using Eqs. (3.3), whereas the flexibility matrix is transformed by

(3.8)

leading to the same matrix as given by Eq. (2.25).

From now on, if no other reference is made, internal forces and deformations are assumed rel-

ative to the local axis systems above introduced.

For convenience purposes in later developments, the following decomposition of and is

introduced

Ss x( ) Qes

Ss x( ) bes x( ) Qes⋅ Spsx( )+=

bes x( ) bs x( ) Tes⋅= bes x( )

bes x( )

1 0 0 0 0 00 0 0 1 L⁄ 0 1– L⁄0 0 1– L⁄ 0 1 L⁄ 00 1 0 0 0 00 0 1 x L⁄–( ) 0 x L⁄ 00 0 0 1 x L⁄–( ) 0 x L⁄

=

ues besT x( ) es x( )⋅ xd

0

L

∫=

F TesT Fes Tes⋅ ⋅= with Fes bes

T x( ) f x( ) bes x( )⋅ ⋅ xd0

L

∫=

u ues

Page 94: Seismic Assessment of Reinforced Concrete Frame Structures

70 Chapter 3

(3.9)

where each partial contribution has non-zero value only for one type of displacement compo-

nent; for these contributions are given as follows

(3.10)

Similar expressions apply to and an identical decomposition can be applied to and .

3.3 Trilinear model

In the present study a global section model has been adopted to reproduce the non-linear

moment-curvature constitutive relation. It is mainly based on an existing trilinear-type model

(Kunnath et al. (1992)), consisting on a primary or skeleton curve, shown in Figure 3.2, com-

bined with three control parameters for the hysteretic rules in each bending direction. It

appears to be a quite versatile Takeda-type model, since a wide variety of hysteretic features

can be reproduced with an adequate choice of the trilinear curve and the control parameters.

The skeleton curve is defined by typical turning points associated with characteristic stages of

the RC section behaviour, namely the cracking (C) and yielding (Y) points, and a post-yielding

branch. The post-yielding stiffness can be indirectly obtained by establishing a ultimate state

point (U) associated to a pre-defined section limit state. All these entities can be different for

positive and negative bending directions indicated by the corresponding superscripts.

The turning points, as well as the ultimate state point, can be identified by means of well

known cracking, yielding and ultimate state criteria along with a fibre-type section discretiza-

tion, thus accounting for the behaviour of steel and concrete through adequate material models.

In this method the section is divided into fibres or layers, to which the specific material models

are assigned, so as to reproduce their uniaxial behaviour. Such formulation is implemented in

u uux u

θx uθy u

θz+ + +=

ues uesux ues

θx uesθy ues

θz+ + +=

u

uux

ux1

00000⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= uθx

0θx1

0000⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= uθy

00θy1

0θy2

0⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= uθz

000θz1

0θz2⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

ues Q Qes

Page 95: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 71

the general purpose computer code CASTEM2000 (CEA (1990)), and details can be found in

Guedes et al. (1994).

A general fibre-based procedure can be established for all kind of sections, but, since rectangu-

lar and T-shaped sections are of particular interest, a specific algorithm has been written, based

on analytical expressions. Therefore, a very efficient procedure has been developed (Arêde and

Pinto (1996)) and implemented in CASTEM2000, allowing to account for the confinement of

the concrete core and the strain hardening of steel. Since this corresponds to a pre-processing

task, at the level of input data preparation, the description of that procedure is presented in the

context of implementation notes in 4.2.

Figure 3.2 Primary or skeleton trilinear curve for the global section model.

The hysteretic behaviour is controlled by a set of rules governing the loading, unloading and

reloading phases. Loading is considered when the state point moves along the primary curve,

while unloading occurs for load intensity decreasing without sign inversion. Reloading is asso-

ciated with increasing load intensity, generally after sign inversion and before reaching again

the primary curve. Hysteretic rules can account for the following phenomena:

• unloading stiffness degradation for increasing inelastic deformations;

• the pinching effect, related with the reduction of the hysteresis loops area, due to pro-

nounced decrease of reloading stiffness followed by some stiffness recovering beyond a

certain threshold of plastic deformation;

My+

Mc-

My-

Mc+

ϕ

M

ϕc+

ϕc-ϕy

-k0

1kc

+

kp-

kc-

kp+

ϕy+

C+

Y+

C-

Y-

O

Page 96: Seismic Assessment of Reinforced Concrete Frame Structures

72 Chapter 3

• strength degradation due to deterioration of concrete resisting capacity, as a consequence of

cyclic loading along with increasing plastic deformation.

An exhaustive model description is deemed unnecessary in this work, since the model is in

existance and has been implemented and tested (Kunnath et al. (1990), Kunnath et al. (1992));

however, some details are included in Appendix B, in order to clarify its most relevant fea-

tures, capabilities and limitations.

This model has been chosen because it is quite versatile in reproducing several familiar phe-

nomena of RC section behaviour; nevertheless, other models could be adopted with different

rules for stiffness and strength degradation or the pinching effect, since they do not signifi-

cantly affect the development strategy underlying the global element model.

3.4 Control sections and element zones

3.4.1 Definitions

The control sections, involved in the flexibility element state determination, are introduced in

this chapter. Between control sections, distinct element zones are defined in order to perform

integration for element flexibility and displacement residual computation.

The element is first assumed divided into two parts where the section properties are considered

uniform and equal to those of the corresponding end section; this assumption requires a span

control section where the transition of properties is considered. It is also assumed that yielding

can occur only in zones adjacent to the end sections, whereas cracking is admitted at any span

section. Therefore, in cases where plastic sections may be expected along the span, a subdivi-

sion into more elements should be adopted.

Two groups of control sections are considered:

• Fixed sections: this group includes the element end sections E1 and E2 and a span section H,

dividing the element into the two parts mentioned above and shown in Figure 3.3. End sec-

tions are fully controlled by the adopted global section model; therefore, the corresponding

cracking and yielding stages, as well as pinching and degradation of stiffness or strength

can be taken into account. The span section H is controlled by a simplified model (see 3.5)

allowing for the cracking stage to be considered; since it is a transition section, a double

Page 97: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 73

control is performed because distinct properties may be assigned to its left and right sides.

This ensures that sections belonging to one specific part of the element are governed

according to the adequate properties and model.

Figure 3.3 Element control sections: fixed sections (E1, E2 and H) and moving section (M)

• Moving sections: this section type is associated to special points in the moment-curvature

relation (from now on referred to as M-ϕ curve), which correspond to the cracking ,

yielding and null moment (O) points, as shown in Figure 3.2. Since they have fixed

moment values, the corresponding section position along the element has to change when-

ever the moment distribution is modified. Actually, the movement of these sections is only

possible because their points in the M-ϕ curve are known and univocally defined.

The treatment of fixed sections is quite normal, requiring only the respective model to be

defined. By contrast, the positions of moving sections need to be found, prior to their state

determination, which can be readily achieved once the moment distribution is known.

Denoting by the moment for a specific moving section, then its position is associated with

the abscissa (see Figure 3.3, where the usual representation of positive moments down-

wards is adopted) which can be obtained from

(3.11)

where the left-hand term is the moment distribution, written in terms of end section moments

and , relative to the respective axis systems as defined in Figure 3.1, and of the ele-

ment applied loads p generically referring to distributed or concentrated loads.

E1

M*

E2Hx*

+- M x( )

xM

sH1sH2

s1 s2

C+ C-( , )

Y+ Y-( , )

M*

x*

M x* p ME1ME2,, ,( ) M*=

ME1ME2

Page 98: Seismic Assessment of Reinforced Concrete Frame Structures

74 Chapter 3

For convenience, local abscissas ( and ) are also introduced (see Figure 3.3) in association

with each end section or element part ( for the left part and for the right

one), as this helps to clarify the scheme for determination of internal moving sections.

Since these sections will be always labelled relating to the index i of the element part they

belong to, the subscript i in is suppressed whenever that abscissa refers to a specific internal

moving section; similarly, for the span section H, refers to its local abscissa in the element

part . According to this convention, Eq. (3.11) can be also written in terms of local abscis-

sas

(3.12)

and, depending on several factors, it may have none, one or more than one solution, as dis-

cussed later on. In any case, it represents the general procedure to obtain moving section posi-

tions, whichever they are, and its solutions are included in Appendix C, for the cases of linear

and parabolic moment distributions, when a given moment is imposed. Due to the impor-

tance of moving sections, they are further detailed in the following sections, separately accord-

ing to their nature.

3.4.2 Cracking sections

Cracking sections are associated to each end section and can be obtained through Eq. (3.12)

by imposing , where stands for the positive or negative cracking moment at

the end section . The following assumptions are considered and commented later in 3.5.3.1:

• once a section has reached cracking, it will not revert back the uncracked state; therefore,

these sections can move along the element in a unique direction;

• if cracking occurs for a certain direction of bending, then the section is also assumed in

cracking conditions for the opposite direction.

Along with these assumptions, the concepts of cracking and cracked sections must be distin-

guished: in a cracking section, the cracking moment has been reached but never exceeded,

whereas in a cracked section that threshold has been previously overcome.

Therefore, depending on the evolution of the bending moment diagram from one step to

another, a cracking section can either remain as a cracking one, or evolve to a cracked section.

s1 s2

Ei EiH i 1= i 2=

si

sHi

EiH

M si* p ME1

ME2,, ,( ) M*=

M*

Ei

M* Mci

+/-= Mci

+/-

Ei

Page 99: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 75

Figure 3.4 illustrates these possible situations referring to the evolution of a cracking section

, generated by the moment diagram at the k-th load step.

Figure 3.4 Distinction and evolution between cracking and cracked sections

If this diagram evolves to one of those indicated by for the step k+1, the cracking

section turns to a cracked one, and the new cracking section shifts to the right side (in cor-

respondence with the small circles). On the contrary, if is the new diagram,

remains as the cracking section.

The number of cracking sections likely to develop depends on the element applied loads and

must be discussed accordingly.

For the case of no element loads (linear bending moment diagram in the whole element), or for

a concentrated force applied on section H (linear distribution of moments in each element

part), at most two cracking sections can develop along each element part, as shown in Figures

3.5-a) and b).

These sections are considered of two types:

• outermost or extreme cracking sections ( or ), which monotonically move apart from

the corresponding end section ( or ) and are associated with bending moment diagram

decreasing (in absolute value) from the end section towards H or the null-moment section;

• innermost or span cracking sections ( or ), which, by contrast, move towards the

respective end sections and correspond to a moment diagram increasing (in absolute value)

from the null-moment section towards H.

It must be noticed that, in case null-moment sections do not exist inside the element, the above

definitions are still valid if reference to null-moment sections is replaced by reference to end

sections.

Ck Mk x( )

Mc1

-

+

-

Mk x( )

Mc1

+ Ck

Mk+1 x( )

Mk+1 x( )Mk+1 x( )′

Ck evolves to cracked section

Ck remains cracking section

Mk 1+ x( )

Ck

Mk 1+ x( )′ Ck

C1 C2

E1 E2

C3 C4

Page 100: Seismic Assessment of Reinforced Concrete Frame Structures

76 Chapter 3

Figure 3.5 General layout of assumed cracking sections and local abscissas for no element

loads or concentrated force applied in H

For each element part , the extreme and span cracking sections are simply denoted by

and and the corresponding local abscissas are and .

In the case of uniformly distributed force, either with or without a concentrated force applied

on section H simultaneously, the bending moment distribution is parabolic and a maximum of

four cracking sections can be defined in each element part as shown in Figure 3.6.

Actually, the location where cracking starts in this case cannot be easily anticipated due to the

superposition of moments generated by end section loads and by the distributed force. Depend-

E1

Mc1

-

E2H

sc1

+- M x( )

x

Mc1

+ Mc2

+

Mc2

-

C1 C2C3 C4

sc2

Uncracked

Crackeds1 s2

sH1sH2

E1

Mc1

-

E2H

sc1

+- M x( )

x

Mc1

+ Mc2

+

Mc2

-

C1 C2C3 C4

sc3sc2

sc4

Uncracked

Cracked

a) No element loads

b) Concentrated force applied in H

EiH Ci

Ci+2 sCisCi 2+

Page 101: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 77

ing on the relative importance of these two loading types, cracking does not necessarily start at

the end sections Ei or the span section H; instead it may develop first at any span section where

the maximum moment is found. Therefore, two other cracking sections have to be introduced

( and ) in order to adequately bound the uncracked and cracked zones likely to

develop.

Figure 3.6 Cracking sections and local abscissas for parabolic moment distribution.

However, the total of eight cracking sections in the whole element can hardly develop simulta-

neously. Since cracking moments are usually almost uniform along the element, the full devel-

opment of the four sections in one element part (as assumed in Figure 3.6) would induce full

cracking of the other part, for which only the respective end section and the section H should

be monitored. Furthermore, such a moment diagram is very unlikely in the context of structural

response under cyclic lateral loads.

Sections and are activated only when the maximum moment section does not occur

in the fixed control sections (i.e. when vanishes in one of the element parts), which

is more expectable for low intensity of lateral loads (thus, less relevant end section moments),

precisely when cracking often initiates and is of more importance. On the other hand, these

sections are mainly relevant to define the cracked zones in the internal part of the element,

which are less important for the structural response to lateral loads; additionally, in the context

of cyclic loading and due to the unique moving direction rule of cracking sections, and

can be overridden by the development of and in subsequent steps of loading.

Ci+4 Ci+6

sHi

Ei

Mci

-

H

sci

M si( )Mci

+

Ci Ci+6Ci+2 Ci+4

sci 2+

Uncracked

+

-

Crackedsci 4+

sci 6+

MEi

MH

si

Ci+4 Ci+6

d M dsi⁄

Ci+4

Ci+6 Ci Ci+2

Page 102: Seismic Assessment of Reinforced Concrete Frame Structures

78 Chapter 3

If lateral load effects are predominant over those generated by the distributed force p during

the cracking initiation phase, the sign of tends to be uniform in each element part

(thus for maximum moment occurring in the fixed control sections) and cracked zones can be

adequately bounded by sections and , only.

For example, in the case of a cantilever beam subjected to a distributed load p and a tip force F,

schematically illustrated in Figures 3.7-a), sections and cannot develop along the

element; only cracking sections and are allowed to appear, and the latter only for the

step k´. Also, in the case of a simply supported beam shown in Figures 3.7-b), under the action

of alternating end section applied moments and a constant distributed force p adequately

selected to prevent the development of the maximum moment in the span, sections and

are sufficient to control the cracked zones likely to form. It is worth mentioning that, the

algorithm presented herein has been applied in these two cases, for a distributed force p, and it

has performed quite well under the restrictive assumption of sections and only.

Obviously, the relative predominance of lateral loads or distributed force cannot be easily

anticipated and the development of cracking sections and should be taken into

account. However, from the implementation standpoint, the control of such sections was found

quite cumbersome for the present stage of algorithm development and likely to lead to a very

heavy numerical scheme. In fact, for this yet unexplored modelling strategy, a very accurate

description of cracking zones in the central part of the element was deemed of lower priority,

because:

• It stands only for a temporary stage of the loading history and occurs in element locations

less relevant for the response to lateral cyclic loads.

• It is also affected by several other assumptions such as: the fact that cracking in one bending

direction induces automatically cracked behaviour in the opposite direction, which for the

central zone may not be true due to the local predominance of element applied loads; the

consideration of cracking moments and post-cracking stiffness involving only the section

behaviour, without taking into account effects of tension stiffening; the uncertainty related

with the collaborating slab width to be considered in the section modelling; the approxima-

tion of distributed force by a uniform diagram, which for very common cases is quite debat-

able (as for example when bi-directionally reinforced concrete slabs lead to force

distributions in the beams better described by a bi-triangular diagram, rather than a uniform

one).

d M dsi⁄

Ci Ci+2

Ci+4 Ci+6

C1 C4

Ci

Ci+2

Ci Ci+2

Ci+4 Ci+6

Page 103: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 79

Figure 3.7 Examples of restricted cracking sections in the presence of distributed force

Despite a first implementation phase in which the uniformly distributed force p was effectively

considered along with the control of sections and only, in order to assess the algorithm

ability to treat simple and adequately chosen non-critical cases (as those referred above and

shown in Figures 3.7), the aforementioned reasons justify the adoption of a simplified strategy

for considering the effects of p without developing sections and .

The approximate procedure consists in replacing the uniformly distributed force p by a concen-

trated and equivalent one Peq applied in span section H, usually located at or near the half-

span. Several equivalence criteria can be considered, but the most logical appears to be the one

based on equality between elastic fixed-end moments due to p and those due to Peq. Note that

p is usually the first load to be applied (normally arising from dead and live loads), when the

structure behaviour is elastic or quasi-elastic. With this approximation, the control of cracking

sections reduces to that already described for the concentrated force case shown in Figure 3.5-

b).

Mc1

C1

a) Cantilever beam

E1 E2H

M x( )

ME1

b) Simply supported beam

p

F

ME1ME2

Mc1

C1

E1 E2H

M x( )'

ME2'

Mc2

C4

Step k

Step k´>k

p

Mc1

- Mc2

-

Mc2

+Mc1

+

C1

E1 E2H

M x( )Step k

C2

C4

Mc1

- Mc2

-

Mc2

+Mc1

+

C1

E1 E2H

M x( )

Step k´>k

C3 Full cracked

Ci Ci+2

Ci+4 Ci+6

Page 104: Seismic Assessment of Reinforced Concrete Frame Structures

80 Chapter 3

It is important to note that, from the implementation standpoint, this simplification is quite sig-

nificant due to the particular features of the model adaptations to be referred later (see 3.5). In

fact, such features (particularly the cracking curvature discontinuity), along with the moving

character of cracking sections and the possibility of full cracking in one or both element parts,

require special care for the control of cracking sections and their contributions to the element

flexibility (see 3.6.2); therefore, the more cracking sections are involved, the more elaborated

is the procedure to control them all.

However, in a future phase of this element model development, the consideration of parabolic

moment diagrams and the inherent control of sections and , shall be included in

order to render the algorithm more general. For this reason, the formulation described in this

study already explicitly includes the force p, both in the derived expressions and in some fig-

ures, although the implemented algorithm corresponds to the simplified version above men-

tioned. Therefore, the expressions effectively used for the cases where distributed load exists

in the element, correspond to and .

It is noteworthy that, in each element part, cracking section locations are bounded by the corre-

sponding end section and the span section H; therefore, in case they are found out of the

respective element part, they are assumed coincident with the closest bound and removed from

the control process.

In order to comply with the definition of cracking sections as stated above, an adequate search-

ing scheme is enforced independently for each element part , using the respective local

abscissa (thus, expressing the moment distribution as ). Such scheme consists of the

following steps:

• First, the adequate cracking moment directions are set according to Figure 3.6 and to the

following definitions of and as described in Figure 3.8:

is the sign of cracking moment allowing to obtain (and possibly )

is the sign of allowing to obtain (and possibly ).

• The local abscissas of cracking sections are then obtained from the defining equations and

according to some restrictive conditions summarized in Table 3.1, where the superscripts

and in take the signs as defined above.

The solutions of the defining equations are detailed in Appendix C and the restrictive condi-

Ci+4 Ci+6

p 0= P Peq=

Ei

EiH

si M si( )

b b

b MciCi Ci 6+

b b–= MciCi 2+ Ci 4+

b

b Mci

Page 105: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 81

tions are introduced to establish whether or not a given cracking section can be defined for the

current moment distribution (e.g., the abscissa of section is provided by the solution(s) of

but it can only be considered as a valid solution for if is

positive and if holds in case the section has been previously activated).

Figure 3.8 Definition of cracking moment directions

Table 3.1 Definition of cracking section abscissas

Section Defining Equation Restrictive Conditions

Ci 2+

M si( ) Mci

b= sCi 2+d M dsi⁄( )

Ci 2+

sCi+2sCi+6

≤ Ci 6+

b sign MEi( )= if d M

dsi-----------

Ei

0<

b sign– MEi( )= if d M

dsi-----------

Ei

0>

Mci

- M si( )

+

-

MEi

EiMci

+ +

-

MEi

Ei

Mci

+

-

MEi

Ei

Mci

-

+MEi

Ei

M si( )

M si( )

M si( )

si si

Ci M si( ) Mci

b= d Mdsi

-----------Ci

0<

Ci+6 M si( ) Mci

b= d Mdsi

-----------Ci 6+

0>

Ci+2 M si( ) Mci

b= d Mdsi

-----------Ci 2+

0> sCi+2sCi+6

Ci+4 M si( ) Mci

b= d Mdsi

-----------Ci 4+

0< sCi+4sCi

Page 106: Seismic Assessment of Reinforced Concrete Frame Structures

82 Chapter 3

The consideration of cracking sections, leads to a set of cracked and uncracked zones with a

general configuration as the one shown in Figure 3.6. Thus, for a complete development of

these zones, the following classification holds in each element part:

Obviously, when sections and are not considered, the third and fourth zones do not

exist and is replaced by in the fifth zone.

3.4.3 Yielding sections

Yielding sections are obtained in the same way as the cracking sections, but the process

becomes simpler since yielding is not admitted at span sections. Thus, yielding sections,

denoted by appear associated to the end sections , when is imposed in Eqs.

(3.11) or (3.12); as for the cracking sections, refers to the positive or negative yielding

moment at the end section .

The general layout of the assumed yielding and cracking sections is, as shown in Figure 3.9,

almost analogous to that of Figure 3.5-b). As for cracking sections, the following assumptions

are considered and discussed later in 3.5.3.1:

• the yielded section behaviour is irreversible, i.e. the section stiffness prior to yielding can-

not be recovered, thus enforcing also yielding sections to move in a unique direction along

the element;

• if yielding takes place for a certain bending direction, then a yielded behaviour is also

enforced for loading in the opposite direction.

Again, a clear distinction between yielding and yielded sections is due as for the cracking sec-

tions: in the former case, the yielding threshold has never been exceeded, while for the latter

the yielding conditions have been previously reached. Positions of yielding sections are found

by means of the same procedure as adopted for the extreme cracking sections, replacing the

cracking moments by the yielding ones; therefore, no further explanation should be required.

Since cracking moments are always lower than yielding ones (at least for normally designed

sections), yielding sections will be always behind cracking ones (referring to local abscissas).

cracked uncracked cracked uncracked cracked

Ei Ci– Ci Ci+2– Ci+2 Ci+4– Ci+4 Ci+6– Ci+6 H–

Ci+4 Ci+6

Ci+6 Ci+2

Yi Ei M* Myi

+/-=

Myi

+/-

Ei

Page 107: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 83

Figure 3.9 General layout of assumed yielding and cracking sections

Therefore, for complete development of cracking and yielding sections under the assumption

of concentrated load applied at H, at most a new zone has to be considered per element part,

leading to an updated set as follows:

3.4.4 Null moment sections

So far, most of the schematic examples illustrated in previous figures have shown null moment

sections lying in uncracked zones. Since these zones behave linear elastically, the control of

such null moment sections is easy from the computational standpoint.

However, the same does not hold true if these moving sections lie on cracked or yielded zones,

since, for positive and negative bending directions, distinct behaviours are likely to appear in

post-cracking sections. Due to the irreversibility of cracking and yielding section progression,

at a certain load step, the null moment sections may fall into a previously cracked or yielded

zone. That is the case illustrated in Figure 3.10, where both bending moment diagrams and the

corresponding cracking sections are included for two consecutive steps of loading (k-1 and k).

For simplicity, neither the yielded zones nor the span cracking section for the right element

part (coincident with H), are included.

yielded cracked uncracked cracked

E1

Mc1

-

E2

H

+

-

M x( )

x

Mc1

+Mc2

+

Mc2

-

C1 C2C3 C4sy2

Uncracked Cracked

My1

-

My2

-

Y2Y1

Yieldedsy1

Ei Yi– Yi Ci– Ci Ci+2– Ci+2 H–

C4

Page 108: Seismic Assessment of Reinforced Concrete Frame Structures

84 Chapter 3

Figure 3.10 Locations of null moment sections

It can be seen that, while the null moment section for the step k-1 is lying in an uncracked zone

, for the step k the null moment section falls in the cracked zone .

Since in this zone, sections with positive moments may exhibit different behaviour from those

in negative bending, the transition section must be controlled in order to correctly account

for stiffness variations appearing there.

As it will be seen in the following sections, the simplifications introduced at the section model-

ling level (see 3.5) suggest that, null moment section control is simple only if it lies in a

cracked zone; by contrast, the control is much more demanding in the yielded zone. Therefore,

in the present work, null moment sections are activated and controlled only when they exist in

cracked zones.

Null moment section locations are obtained by a similar but simpler procedure to that cracking

and yielding sections. Imposing in Eqs. (3.11) or (3.12), solutions can be obtained

and checked if they are located inside element parts. However, this must be carried out only

after cracking and yielding sections have been found, in order to check if null moment sections

have to be controlled or not.

E1

Mc1

-

E2

H

+

-

Mk-1 x( )

x

Mc1

+Mc2

+

Mc2

-

Uncracked

Cracked

C1k-1

C2k

C2k-1

Mk x( )

C3k

C3k-1

C1kOk

Ok-1

Cracked at step k Cracked at step k-1

C1k-1 C– 3

k-1[ ] Ok E1 C– 1k[ ]

Ok

M* 0=

Page 109: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 85

3.5 Behaviour of the control sections

3.5.1 Modified trilinear model

3.5.1.1 General

As aforementioned, the control of internal moving sections is only practicable if their repre-

sentative points in the M-ϕ curve are known without the need for tracing out all the behaviour

history. This fact becomes clearer by considering Figure 3.11, which illustrates examples of

possible evolutions of cracking and yielding section points along the trilinear model curve. For

both sections, it is assumed that at the load step k their points in the curve are and , coin-

ciding with the curve turning points. For subsequent steps k+1, k+2,..., these section points

occupy the positions and (j=1,2,...), corresponding to a sequence of loading, unloading

and reloading.

Figure 3.11 Evolution of cracking and yielding section points in the model diagram

It is apparent that passing from to presents no problem for the section state determina-

tion, since all the necessary data are available in the primary curve (previous point and stiff-

ness, as well as new stiffness). However, once an unloading step occurs ( to ), the new

stiffness becomes dependent on the inversion point, which, furthermore, becomes necessary

also for subsequent reloading steps. This point, although lying on the basic curve, must be

stored for each step and for all cracking sections continuously appearing. The same reasoning

holds for yielding sections.

Obviously, such requirements are practically impossible to be achieved and thus some simpli-

fying assumptions must be included, in order to perform the state determination of internal sec-

tions using only the primary curve data and the section moment (or curvature) for the step

under analysis.

C0 Y0

Cj Yj

ϕ

M

C0

Y0

C1

C2

Y1

Y2

C0 C1

C1 C2

Page 110: Seismic Assessment of Reinforced Concrete Frame Structures

86 Chapter 3

3.5.1.2 Motivation for model modification

The simplifications adopted in this work arise from the recognition of the detailed shape of M-

ϕ curves for RC sections. Actually, the trilinear curve is just a possible approach to the real

one, and its validity depends on several factors related to the characteristics of material model,

the section geometry, the steel contents and, last but not least, the acting axial force.

The real diagram M-ϕ of a given section is very difficult, if not impossible, to obtain by exper-

imental means, since the measurement of curvatures is always performed over a finite length.

Any M-ϕ curve experimentally obtained is an average diagram over a certain zone, rather than

the effective local measure corresponding to a well defined section. Thus, in this context, the

concept of real M-ϕ curve is not strictly meaningful.

Nevertheless, a good approximation to the actual section moment-curvature behaviour can be

achieved using the above referred fibre model (Vaz (1992), Guedes et al. (1994)) for the dis-

cretization and analysis of sections. In the fibre model formulation as implemented in the com-

puter code CASTEM2000, the main features of the adopted concrete and steel models (for

monotonic loading) are very similar to those assumed in the procedure for the M-ϕ primary

curve definition as used in the present study (see 4.2). However, two differences are found that

should be referred:

• the post-yielding behaviour of steel is more refined in the fibre formulation context, since a

yielding “plateau” is considered and a fourth degree polynomial is used for the hardenning

range; however, with slight modifications in model parameters, it can be forced to fit the

bilinear approach used in Arêde and Pinto (1996);

• the concrete tensile behaviour in the fibre formulation includes a linear softening branch

after the tensile strength is reached, whereas in the present study the strength drops immedi-

ately to zero once cracking occurs; again, the two assumptions can be made to fit by ade-

quate choice of model parameters, but the inclusion of concrete tension softening may be

appropriate as discussed later.

The aforementioned fibre formulation allows to obtain the section M-ϕ curves and, in associa-

tion with a linear Timoshenko finite element discretization (Guedes et al. (1994)), the non-lin-

ear behaviour of beam/column structural components can be traced out. On that basis, a

sensitivity study of section M-ϕ curves can be found in Arêde and Pinto (1996), aiming at the

Page 111: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 87

validation of the proposed procedure to obtain the trilinear primary curve. Several RC sections

have been analysed by the fibre model and compared with the trilinear M-ϕ curve, for varying

conditions related to the total amount of longitudinal reinforcing steel, the ratio of compression

to tension steel contents, the slab width participating in beam behaviour and the axial force.

A detailed description of that study is out of the scope of the present work, but, since the axial

force has been recognized as the most important factor affecting the typical shape of M-ϕ

curves, an example of a symmetrically reinforced concrete section with two levels of axial load

is analysed below. The section geometry, the reinforcement and the basic parameters for the

material models are illustrated in Figure 3.12; more details about the concrete model can be

found in 4.2. The axial load level is defined by the normalized axial force ν, expressed in the

usual way as , and has been adopted with the values 0 and 0.1. Figure 3.12

includes also the scheme of a cantilever beam with uniform section, discretized with constant

curvature Timoshenko elements, in order to obtain the force-displacement diagrams for both

section modelling strategies: the fibre formulation and the trilinear curve.

Figure 3.12 Comparison of fibre and trilinear section modelling formulations. Section, member

and model data

ν N bdfc0( )⁄=

b=0.45m

h=0.

45m

d=0.

41m

4 φ 20

4 φ 20

φ 10 @.075

fc

εcεc0

fc0

fctEc

Confined

Unconfinedfcc

0.2*fcc

fc0 = 44.8 MPa

fct = 4.48 MPa

εc0= 2 (*10-3)

Ec = 33.7 GPa

fs

fy

εsεy εu

Esh

Concrete Steelfy= 570 MPaεy= 2.85 (*10-3)

Esh= 0.87 GPa

εu= 100 (*10-3)

L=3.5m

1 5 8 3

N

No. and length of Timoshenko elements:

Cantilever beam

Section

.5 2.0m

a) Section detailing

c) Concrete and steel model data

F

u

.05

b) Member layout and discretization

Page 112: Seismic Assessment of Reinforced Concrete Frame Structures

88 Chapter 3

The M-ϕ curves obtained from the application of both section formulations, as well as the cor-

responding cantilever tip force-displacement diagrams F-u, are compared in Figure 3.13 for

the two levels of axial force.

Figure 3.13 Comparison of fibre and trilinear section modelling formulations. Local and global

response for two axial load levels

The following aspects can be observed at the section level:

• For zero axial force (the usual case of beams) the post-cracking behaviour exhibits a tempo-

rary decrease of resisting moment due to a progressive transfer of stress from cracked con-

ν = 0.0 ν = 0.1

Displ.(m).

Force (kN)

Displ.(m)

Force (kN)

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 X1.E-2

0.

20.

40.

60.

80.

100.

120.

140.

Curv.(m-1)

Mom.(kN.m)

.0 .3 .6 .9 1.2 1.5 1.8 2.1 2.4 2.7 3.0 x1.E-2

0.

.5 1.0 1.5 2.0

2.5 3.0 3.5 4.0 4.5

5.0 x1.E2

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 X1.E-2

0.

10.

20.

30.

40.

50.

60.

70.

80.

90.

Fibre Trilin Tril_mod

a) Moment- curvature diagrams

b) Tip force-displacement curves

ν = 0.0 ν = 0.1

Fibre Trilin

Curv.(m-1)

Mom.(kN.m)

.0 .3 .6 .9 1.2 1.5 1.8 2.1 2.4 2.7 3.0 x1.E-2

.0 .5 1.0 1.5

2.0 2.5 3.0 3.5 4.0

4.5 5.0 x1.E2

Page 113: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 89

crete fibres to uncracked ones and to the steel; after a certain point, the influence of

tensioned steel becomes dominant and the resistance increases again with a stiffness that

tends gradually to the fully-cracked section stiffness (i.e., when all the concrete in the ten-

sioned area is considered inactive). In such case it can be seen that a trilinear curve does not

fit well in the post-cracking zone, where the fully-cracked stiffness (pointing from the ori-

gin to the yielding point) would be more appropriate.

• For the case of , the transition phase due to cracking is not so pronounced because

the acting axial force implies higher compressive forces in the section; therefore, the ten-

sioned concrete area is less important to the section equilibrium and the stress transfer

becomes less visible during the transition phase. The trilinear curve appears now to be a

good approximation.

The force-displacement results agree with the above comments. Actually, for the null axial

force case, the “lack” of curvature shown by the trilinear curve between the cracking and yield-

ing points, leads to significant lower displacements although with tangent stiffness not very

different from that of the fibre formulation.

By contrast, the good fitting of M-ϕ curves when the axial load is present, leads also to an

excellent agreement of force-displacement curves.

It is important to note that, for , the immediate post-cracking strength drop in the M-ϕ

curve is much more pronounced than in the F-u response of the element. In the case of the

results shown, this is just due to the fact that cracking is not occurring simultaneously in all

sections. However, two other phenomena, not considered in the above results, can also contrib-

ute to that fact, namely:

• The inclusion of concrete tension-softening with finite slope, instead of a sudden elimina-

tion of tensile strength whenever the peak tensile strain is exceeded (i.e., an infinite tension-

softening); this aspect is strictly related to the local section behaviour modelling and tends

to increase the cracking moment and to smooth the strength drop (already at the M-ϕ level).

This is shown in Figure 3.14, where the section fibre modelling includes now a linear sof-

tening branch with finite slope for the tension behaviour model, also illustrated in the M-ϕ

diagram. The dashed vertical line in the tension model corresponds to the assumption made

in the case of Figure 3.13. However, the cracked stiffness still fits quite well the M-ϕ curve

in the post-cracking phase.

ν 0.1=

ν 0=

Page 114: Seismic Assessment of Reinforced Concrete Frame Structures

90 Chapter 3

Figure 3.14 Effects of tension-softening in the fibre formulation

• The fact that, in a given element zone, bending moments have exceeded the cracking

moment, does not imply all the sections to be effectively cracked; due to bond between rein-

forcement and the surrounding concrete, gradual redistribution of internal forces take place

from concrete to steel leading to a crack pattern that tends to stabilize at a certain finite

spacing (Feenstra (1993)). Therefore, a cracked zone has a higher stiffness than if consid-

ered with a uniform cracked stiffness as taken from the M-ϕ curve. This phenomenon,

known as tension-stiffening, is related with the behaviour of the global element (or part of

it) rather than with the local section behaviour; thus, its effect appears only in the global

response and tends to smooth even more the post-cracking transition phase.

Feenstra (1993) discusses the superposition of these two effects in RC members, though in the

stress-strain context, which can be also considered in the M-ϕ space for a clear understanding

of the cracking transition phase. However, the tension-stiffening effect is often indirectly mod-

elled by means of tension softening in the concrete model (Figueiras (1983), Póvoas (1991)).

3.5.1.3 Proposed model modifications

In order to partially account for the above described behaviour, a modified M-ϕ curve would

be desirable. On the one hand, the post-cracking strength drop has very localized effect which,

furthermore, is negligible when global element behaviour is sought; on the other hand, it is par-

tially counterbalanced by finite tension-softening of concrete and by the tension-stiffening

effect along the member. It is clear, however, that a transition from the uncracked to the fully-

cracked stiffness has to be included (at least for cases with zero or low axial force level) if the

ν = 0.0

Displ.(m).

Force (kN)

Curv.(m-1)

Mom.(kN.m)

.0 .3 .6 .9 1.2 1.5 1.8 2.1 2.4 2.7 3.0 x1.E-2

.0 .5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 x1.E2

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 x1.E-2

0.

10.

20.

30.

40.

50.

60.

70.

80.

90.

fct

3εctεct εc

TensionModel

Fibre Trilin Tril_mod

Page 115: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 91

post-cracking behaviour is to be adequately traced out. With this in mind, a modified M-ϕ pri-

mary curve is proposed as follows:

• sections behave linearly with elastic stiffness up to the cracking moment;

• at imminent cracking, the transition is enforced from uncracked to fully-cracked stiffness, at

constant moment (the cracking one); this means that discontinuities arise in both stiffness

(or flexibility) and curvature distributions along the member;

• for incipient yielding, sections are likely to change to the post-yielding branch as in the

usual trilinear case, thus inducing discontinuity only in the stiffness distribution.

The corresponding primary curve is also included in the M-ϕ diagrams of Figures 3.13 and

3.14 (under the reference of Tril_mod) and, as already pointed out, an excellent fit with the

fibre formulation curve is obtained when axial force is negligible. If the axial force becomes

significant, the two curves do not agree so well. A possible solution to approximately over-

come this drawback could be the adoption of an intermediate post-cracking stiffness (between

the uncracked and the fully-cracked one), estimated in such a way to compensate the excess of

member displacements caused by the use of the fully-cracked branch. Indeed, this was tried for

some column cases for which the force-displacement response obtained by the proposed ele-

ment model, with an intermediate stiffness, tended to the fibre modelling response. However,

any estimate of such a stiffness depends on the expected level of end section deformation and

the distribution of curvatures along the element. Since this could not provide a general criteria,

it was decided to keep the fully-cracked stiffness for post-cracking behaviour in the present

stage of model development, although bearing in mind that column deformations might

become somewhat overestimated when the behaviour is mostly in the post-cracking range.

The hysteretic rules associated to the modified trilinear M-ϕ curve are shown in Figure 3.15

and can be stated as follows:

• loading, unloading and reloading of uncracked sections take place along the linear elastic

branch (1) with flexibility ;

• for cracked sections, i.e. those for which (or ), where

is the maximum moment (positive or negative) experienced up to the present load

stage, loading occurs along branches 2 (or 3) up to yielding, while unloading and reloading

are done pointing to and from the origin along branches 6 (or 7) with flexibility (or );

this means that sections between the cracking and yielding ones are not allowed to have

residual deformations at zero moment;

f0

Mc+ Mmax My

+≤< My- Mmax Mc

-<≤

Mmax

fy+ fy

--

Page 116: Seismic Assessment of Reinforced Concrete Frame Structures

92 Chapter 3

• the case of yielded sections, i.e. for which (or ), is treated with

almost the same rules of the trilinear model described in Appendix B, namely concerning

unloading and reloading with possible pinching or strength degradation; for subsequent pur-

poses, the following generic notation is adopted for the flexibility of yielded sections:

for loading, for unloading and for reloading.

Two relevant differences, relative to the rules stated in Appendix B, shall be referred:

• for unloading in yielded zones, the common point is now on the lines (or ), in

order to cope with the fact that a given yielding section unloads pointing to the origin; there-

fore, continuity of stiffness distribution exists at the yielding section for unloading condi-

tions, but the same parameter, as defined in Appendix B, leads now to a greater amount of

stiffness degradation;

• after a section has yielded for a given bending direction (e.g. ), but not actually for

the opposite direction (e.g. ), then, for reloading in this opposite direction, the

maximum point is taken at the yielding point ( in Figure B.2) unless the pinching

effect has to be also considered; in fact, this procedure is similar to those adopted in several

other models based on a bilinear primary curve (e.g. Filippou et al. (1992), Coelho (1992)).

Figure 3.15 Hysteretic rules for the modified trilinear model

The modified trilinear model features allow an immediate knowledge of each internal section

Mmax My+> Mmax My

-<

fp+/-

fu+/- fr

+/-

OY+ OY-

α

M My+>

Mmax My->

E- Y-≡

f0

fy+

fp+

Mc-

Mc+

ϕ

M

C+

My+

Mc-

My-

Mc+

ϕ

M

ϕy- ϕy

+

C+

Y+

C-

Y-

Cc+

Cc-

C-

1

1

3

7

2

6

My+

My-

ϕ

M

ϕy- ϕy

+

Y+

Y-

7

6

5

4

fy-

fu+

fp-

fu-

a) Uncracked sections b) Cracked sections c) Yielded sections

o o o

Page 117: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 93

state point in the M-ϕ diagram, without the need for all the previous loading history. Actually,

with the exception of yielded sections (i.e., between end and yielding sections), the state deter-

mination for any internal section is straightforward and can be done in terms of either input

moments or input curvatures. This is obviously a consequence of the one-to-one nature of the

assumed M-ϕ curve and is further detailed in the next section.

Finally it must be underlined that, despite the fact that model modifications have been intro-

duced reporting to internal moving section behaviour, the modified model is adopted also for

the fixed sections, namely the end sections and the span (H) one, because they are likely to

pass by internal state phases similar to those of moving sections. The difference lies only on

their fixed character, which allows for their loading history to be monitored.

3.5.2 Transition from uncracked to cracked section behaviour

The transition from uncracked to fully-cracked behaviour is performed along a constant

moment plateau, herein designated by cracking plateau, which, according to 3.4, is assumed

uniform in each element part ( and ).

If one looks at the behaviour of an isolated section following such a model, this plateau

assumption appears to be a major drawback since it introduces an indetermination when the

cracking moment is reached. However, if the element moment distribution is taken into

account, it becomes apparent that such indetermination can hardly occur in most practical situ-

ations. Referring to a given element part the two following cases can take place:

• Non-uniform bending moment distribution

The cracking moment is found at a finite number of sections where a “jump” in the curva-

ture distribution is assumed, as shown in Figure 3.16 for linear moment distribution.

It can be seen that no state points exist between and in the M-ϕ diagram, regardless of

the way of crossing the plateau; for this reason the plateau is traced as a dashed line.

The left and right sides ( and ) of the cracking section have their state points in the M-

ϕ diagram coincident with and , respectively. The curvature diagram exhibits discon-

tinuity between these two points, which introduces a special contribution in the element

flexibility matrix due to the section that is changing with the applied moments. This

aspect is crucial to successfully pass the cracking plateau and is detailed in 3.6.2 for the cor-

rect flexibility derivation.

E1H E2H

EiH

C Cc

CL CR

Cc C

C

Page 118: Seismic Assessment of Reinforced Concrete Frame Structures

94 Chapter 3

Therefore, if the influence of the cracking plateau is adequately taken into account, it can be

used for the cracking transition with no risk of indetermination because, in this loading

case, no section state point really exists in the plateau.

Figure 3.16 Cracking transition in the case of non-uniform moment distribution

• Uniform bending moment distribution

This loading case, although not so common in practical situations, is more troublesome

when the cracking plateau has to be overtaken. The case of may arise either

due to combinations of end section applied moments and applied force at H exactly match-

ing the moment in that element part, or due to rotations imposed at the end sections

such that the required uniform curvature lies between and , as shown in Figure

3.17. The former case leads to an indeterminate solution, while the solution for the latter is

readily known ( ), but can it be numerically obtained only if a special iterative procedure

is adopted to bypass the zero stiffness problem. It is clear, though, that all sections have

their state points in the cracking plateau; curvature discontinuities do not appear and the

whole element part is controlled by the end section.

In order to handle both situations of moment distribution, a specific algorithm can be adopted

for the cracking plateau transition based on the following:

• Instead of assuming a constant moment plateau, a very small non-zero stiffness is assigned

to the cracking transition; this is meant to avoid the solution indetermination when, for

example, a constant moment is applied.

CCL CR

Left Right

ϕc ϕ si( )

Mc

ϕ

M

C

L

oϕcL

ϕcR

Cc

R

Mc

M si( )

C

∆ϕcc

∆ϕcc

Ei H

si

HEi

M si( ) Mc=

Mc

ϕa( ) ϕc ϕcc

Mc

Mc

Page 119: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 95

• Any section whose state point lies on the cracking plateau is governed by the fully-cracked

flexibility , for loading (i.e, above the plateau) and a secant flexibility or for unload-

ing or reloading below the plateau.

Figure 3.17 Cracking transition in the case of uniform moment distribution

These last procedures are schematically illustrated in Figure 3.18, where, for simplicity, the

assumed non-zero stiffness of the plateau is not shown. The iterative process for searching the

solution across the plateau is identical to the section state determination procedure shown in

Figure 2.10, but the updated flexibility is always . Two possible situations are considered:

the case where convergence is reached for curvature beyond the plateau (Figure 3.18-a)) and

that for convergence occurring inside the plateau (Figure 3.18-b)).

Unloading from the plateau occurs pointing to the origin (see Figure 3.18-c)) and eventual

reloading on the opposite direction is done aiming at a plateau point with a fraction of curva-

ture “jump” equal to the maximum one previously attained . While

this maximum curvature “jump” is not exceeded, the unloading and reloading flexibilities

remain unchanged; further increase of that “jump” implies the flexibility to be re-activated .

Attention must be drawn to the fact that the scheme of Figure 3.18-a) could be present in all

loading cases (uniform ones included), whereas those of Figures 3.18-b) and 3.18-c) exclu-

sively refer to the uniform loading case reaching the cracking moment or unloading from it.

This procedure has been actually implemented in the present work context, but very poor con-

vergence performance was obtained for current loading cases, i.e., typically consisting of non-

uniform bending moment diagrams. Therefore, it was found preferable to distinguish the two

cases of moment distribution and to treat them accordingly:

fy fu fr

ϕcϕ si( ) ϕa=

Mc

ϕ

MC

oϕcc

ϕcc

CcMc

Ei

ϕcc

ϕa

M si( ) Mc=

Ei

Hsi

H

fy

∆ϕ* ∆ϕCc

-⁄( ) ∆ϕE+ ∆ϕCc

+⁄( )

fy

Page 120: Seismic Assessment of Reinforced Concrete Frame Structures

96 Chapter 3

• If a uniform moment distribution is found for a given element part, the scheme shown in

Figure 3.18 is adopted for the cracking plateau transition; it is noteworthy that, in such a

case, no internal moving sections need to be activated because the behaviour of the whole

element part is controlled by the respective end section.

• Conversely, for the most usual case of non-uniform moment distribution, the direct algo-

rithm is used, where the total curvature “jump” is considered in all the cracked zone sec-

tions for the cracking plateau transition (see Figure 3.16).

Figure 3.18 Rules for progressive transition of the cracking plateau transition

This strategy avoids penalizing convergence performance in the large majority of cases, but

requires extra implementation effort since the following issues have to be taken into account:

• Cases of bending moment distribution close to uniform (low rate of variation along the ele-

ment), may lead to sudden cracking along rather large zones, thus developing significant

displacement residuals to be eliminated; as this may cause non-convergence problems (typ-

ically characterized by iterative solutions “jumping” below and above the correct solution,

but never being able to reach it), numerical solution guiding schemes have to be adopted,

such as the line search technique (Simons and Powell (1982), Criesfield (1982), Marques

(1986)) in order to successfully achieve convergence in the element iterative scheme.

• Both procedures to overcome the cracking plateau (Figures 3.16 and 3.18) have to be made

compatible in the same algorithm.

Mc

ϕ

M

C

oϕcc

ϕc

Cc

fy

f0

fy

Mc+

ϕ

M

C

oϕcc

+ϕc+

Cc

fy+

f0

Mc

ϕ

M

C

oϕcc

ϕc

Cc

fy

f0

fy

b) Convergence INSIDE the plateau

fy-

fu+

fr-

∆ϕE+

E

∆ϕ*

c) Unloading from or

fr- fu

+ f0–

fy+ f0–

---------------⎝ ⎠⎜ ⎟⎛ ⎞

fy- f0–( )=fu

+ ϕE+

Mc+

--------= =>

a) Convergence BEYOND the plateau

reloading to the plateau

∆ϕCc

+

∆ϕCc

-

Page 121: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 97

3.5.3 State evolution of control sections

3.5.3.1 Internal moving sections

The fact that internal moving sections (cracking, yielding and null-moment sections) are asso-

ciated to turning points in the model curve, where flexibility and/or curvature discontinuities

may occur, implies the left and right sides of each section to be monitored independently. For

this purpose, an adequate convention must be adopted for the section side identification; since

internal sections are associated to end-ones, it is assumed that the left side of an internal section

is the side existing in the interval between that section and the corresponding end-one, while

the right side is the one existing outside that interval.

With this convention and recalling the concepts of cracking and cracked section, as well as of

yielding and yielded sections, it follows that:

• a cracking section has one uncracked side and another side behaving as cracked;

• a yielding section presents cracked behaviour in one side and yielded in the other.

This distinction is important because each section side follows different paths for subsequent

loading evolutions. Consider the step for the k-th external load increment; at the beginning of

that step, cracking and yielding sections refer to the previous step k-1 and are denoted by

and , respectively, while at the end, the corresponding updated sections are and .

Depending on the local effect of the load increment (loading, unloading or reloading), the fol-

lowing state evolution cases are likely to appear (referring to Figure 3.15). For cracking sec-

tions, and, for simplicity, assuming the load increment small enough to avoid “jumping”

directly from uncracked to yielded behaviour, the following situations hold:

• unloading or reloading without generation of new (=> ):

the uncracked side remains elastic (branch 1) and

the cracked side follows the lines (branches 6 or 7);

• loading or reloading with generation of new (=> ):

the uncracked side proceeds from elastic (branch 1) to cracked (branches 2 or 3) and

the cracked side remains along the lines (branches 2 or 3);

Ck-1

Yk-1 Ck Yk

Ck-1 Ck Ck Ck-1≡

OY+/-

Ck-1 Ck Ck Ck-1≠

OY+/-

Page 122: Seismic Assessment of Reinforced Concrete Frame Structures

98 Chapter 3

• (just generated):

the uncracked side remains elastic (branch 1) and matches point , whereas

the cracked side moves from uncracked behaviour to the lines (branches 2 or 3),

matching point ;

In turn, for yielding sections the following cases are considered:

• unloading or reloading without generation of new (=> ):

both the cracked and the yielded side follow the lines (branches 6 or 7);

• loading or reloading with generation of new (=> ):

the cracked side proceeds from lines (branches 2 or 3) to yielded behaviour

(branches 4 or 5) and

the yielded side remains along the post-yielding lines (branches 4 or 5);

• (just generated):

both the cracked and the yielded sides match point , being the cracked side on lines

(branches 2 or 3) and the yielded one on the post-yielding lines (branches 4 or 5).

It will be shown later that only the current step (k) configuration of internal moving sections is

of interest for the state determination and for residual computation. Therefore, the behaviour of

the step k-1 moving sections is not needed to be controlled, which renders the numerical imple-

mentation easier and clearer.

For the load step under analysis, all the cracked sections exist between the cracking and the

yielding or the span (H) sections (see Figure 3.9). In any case, their state points lie on the

lines, likely to have different slopes in the positive and the negative direction. Since

these lines point to the origin, the element stiffness distribution may change exactly at the null-

moment section (O), if it happens to fall in a cracked zone. As already mentioned in 3.4.4, if

the section O lies on a uncracked zone its control is not required because the elastic stiffness

does not change with the bending direction. Conversely, if it shows up in a yielded zone, it can-

not be easily controlled since it is a moving section without monitored loading path. Therefore,

the null-moment section is controlled as a cracked section, following the model labelled

cracked in Figure 3.15-b), and is only activated if falling in a cracked zone.

Ck

C+/-

OY+/-

Cc+/-

Yk-1 Yk Yk Yk-1≡

OY+/-

Yk-1 Yk Yk Yk-1≠

OY+/-

Yk

Y+/-

OY+/-

OY+/-

Page 123: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 99

The yielded sections, i.e. those between the yielding and the end sections, appear to be the

most problematic to control, since no simplified and clear model is assigned to them and their

loading history is impracticable to be known. Therefore, as specifically discussed later, yielded

sections will be approximately controlled according to the behaviour of the associated end and

yielding sections simultaneously.

At this point some comments are pertinent about the assumptions concerning the cracking and

yielding development as stated in 3.4.2 and 3.4.3, respectively.

The irreversibility of cracking is a very acceptable assumption in view of the impossibility of

recovering the loss of concrete strength and the inherent stiffness decrease and permanent

deformations. However, for sections in the pre-yielding range and having cracked only for one

bending direction, some temporary stiffness recovery can be observed until cracking occurs for

the opposite bending direction (this is apparent in the results shown in Figure 2.2); after that,

the stiffness progressively decreases towards the fully-cracked stage. This effect, as well as the

permanent deformations actually developing in the pre-yielding range, cannot be taken into

account by the adopted model due to both the irreversibility assumption and the origin-oriented

stiffness considered after cracking.

The temporary recovery of post-cracking stiffness is not of major importance in the cyclic

loading context, because cracking load reversals are very likely to occur in the zones where

stiffness variations are more relevant for the global element behaviour; this sustains both the

irreversibility issue and the assumption that a given cracking section for one bending direction,

is also a cracking one for the opposite direction. In turn, the non consideration of permanent

(residual) deformations leads to the underestimation of energy dissipation in the pre-yielding

range. Despite these shortcomings, the referred assumptions were still considered in order to

assure one-to-one model diagrams that make feasible the control of cracking sections; how-

ever, it is recognized that future improvements should be made on the energy dissipation issue.

Concerning the yielding irreversibility, it is also apparent that, upon load reversals, the pre-

yielding (fully-cracked) stiffness cannot be recovered (again as evidenced in Figure 2.2).

Moreover, if yielding sections were allowed to move back, it would require the control of a

range of sections in the post-yielding branch, which, as referred above, is practically not

achievable. A similar reasoning sustains the need to assume that a yielding section for one

Page 124: Seismic Assessment of Reinforced Concrete Frame Structures

100 Chapter 3

bending direction behaves also as a yielding one for the opposite direction.

3.5.3.2 Fixed sections

The one-side end sections ( and ) are controlled, as explained in 3.5.1, by means of the

curves shown in Figure 3.15, complemented with the cracking plateau transition shown in Fig-

ure 3.18 for the uniform bending moment distributions. After yielding, the hysteretic rules

described in Appendix B apply with two slight modifications referred to in 3.5.1.

The span section H is controlled as the end sections are, both in its left and right sides with the

appropriate model and properties, but the behaviour is limited to uncracked and cracked stages.

However, for uniform bending moment distributions, the explicit control of section H is not

performed because it is taken into account by the behaviour of end section(s).

3.6 Element state determination

3.6.1 General

The general procedure for the element state determination has been outlined and illustrated in

2.4.4, for fixed sections fully controlled by a general model, the basic goal being the calcula-

tion of the total restoring forces (or just the increment ) and the updated stiffness

matrix for a given increment of displacements in the Newton-Raphson iteration n.

According to the notation used in 2.4.4 and the expressions included in Figure 2.10, the whole

process is essentially based on the condensed form expression

(3.13)

where the applied displacements and the residual ones are given by

(3.14)

and the flexibility matrix is obtained according to

E1 E2

Qn ∆Qn

Kn ∆un

∆Qj Fj-1[ ]1–

∆uj ∆urj-1–( )⋅=

∆u…( ) ∆ur…( )

∆uj=1 ∆un= ; ∆ur0 0=

∆uj>1 0 = ; ∆urj-1 bT x( ) ∆er

j-1 x( )⋅ xd0

L

∫=

Page 125: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 101

(3.15)

The final outcome is thus expressed as

(3.16)

where the increment of restoring forces is given by

(3.17)

The major tasks of the element state determination in each iteration j consist on updating the

flexibility distribution and the respective matrix , and on computing residual displace-

ments , in order to check convergence and to set up the next iteration. Both issues are

addressed in the following sections, accounting for the specific features of the element.

3.6.2 Flexibility distribution within the element

The definition of the section flexibility distribution along the element is a key issue of the flex-

ibility formulation, namely aiming at:

• obtaining the element tangent flexibility matrix for the current load step,

• the computation of deformation residuals in the current step and

• the section state determination in the subsequent load step.

If the non-linear behaviour refers only to some of the internal force and deformation compo-

nents, the flexibility distribution needs to be established only for such components since the

other contributions remain constant for the whole loading process. That is exactly the present

case, in which only one bending direction (see 3.1) contributes non-linearly for the flexibility

matrix; the remaining constant contributions are included in Appendix A and just need to be

added to the total flexibility matrix of the element.

The state determination is assumed to be already performed for all control sections, which

means that each moving section location is known, all the uncracked, cracked and yielded

Fj-1 bT x( ) fj-1 x( ) b x( )⋅⋅ xd0

L

∫=

Qn Qn-1 ∆Qn+= ; Kn Fj=converg[ ]1–

=

∆Qn ∆Qj

j 1=

converg

∑=

fj x( ) Fj

∆urj

Page 126: Seismic Assessment of Reinforced Concrete Frame Structures

102 Chapter 3

zones are defined by the adequate internal sections, and each control section flexibility is

updated according to the respective models.

The flexibility distribution is then defined in each zone by linear functions between the bound-

ing section sides lying inside the zone, as shown in Figure 3.19 where two examples of flexi-

bility distributions are included. Figure 3.19-a) refers to the development of yielded, cracked

and uncracked zones associated with and resulting from increasing nodal and element

applied loads. In turn, Figure 3.19-b) shows a typical unloading situation where the three types

of zones are also present, although developed by a previous linear moment distribution ;

for the current distribution none of the moving sections develops further and it is appar-

ent that all the control sections are actually unloading.

Figure 3.19 Examples of flexibility distributions for loading and unloading cases

According to the evolution of the control section state described in 3.5, the possible flexibility

diagrams along the element consist of:

• uniform distributions in uncracked and cracked zones, as a result of the model assumptions;

• uniform distributions in yielded zones, when end sections are loading and yielded zones are

developing (case of Figure 3.19-a));

• linear approximations to the yielded zone flexibility distributions, when end sections are

unloading or reloading and yielding sections are not moving (case of Figure 3.19-b)).

M x( )

M0 x( )

M x( )

Mc1

- Mc2

-

Mc2

+Mc1

+ HC1

C4C3

a) Loading bilinear moment diagram b) Unloading linear moment diagram

C2

E1 E2Y1 Y2

My2

-

My1

-

fy1

- fy2

-fy1

+fy2

+

f01 f02

fp2

-fp1

-

M x( )

f x( )

HC1

C2E1 E2

Y1

Y2

M0 x( )

M x( )

fy1

- fy2

+

f01 f02fu2

+fu1

-

f x( )fp1

-

fp2

+

Page 127: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 103

The flexibility distributions illustrated in Figure 3.19 cover the possible cases for uncracked

and cracked zones. By contrast, yielded zones may present specific problems demanding spe-

cial care, as is the case of Figure 3.19-b) where, for the sharp transition from a uniform to a lin-

ear (non-uniform) distribution when unloading occurs in a yielded zone (see the dashed and

solid line diagrams of ), recourse is made to an event-to-event scheme along with addi-

tional procedures detailed later in 3.6.6.

3.6.3 Element flexibility matrix

Once the flexibility distribution is defined, Eq. (3.8) could be used to obtain the element flexi-

bility matrix . Since only the y bending component is of interest, attention will just focus in

the corresponding non-zero matrix terms given by

(3.18)

is the bending flexibility distribution (as shown in Figure 3.19) and are the force

shape functions contributing to that bending component. These functions are labelled with sub-

scripts l and m corresponding to the end sections and can be extracted from Eq. (3.6) to be sim-

ply written as

(3.19)

The relation between end section indices (l and m) and the corresponding degrees of freedom

(l´ and m´) in the matrix is also included in Eq. (3.18), where the superscript M is labelling

the flexibility terms to highlight that only the bending deformation contribution is considered.

Actually, also the shear distortion contributes to those terms but, due to its linear behaviour, it

has not to be considered explicitly here.

Since is step-wise linearly distributed, the integrals of Eq. (3.18) are divided by zones

where that function is uniquely defined, and that expression is written as

f x( )

Fes

Fl'm'M Flm

M φl x( )fM x( )φm x( ) xd0

L

∫= = for l' 3 5,=( ) m' 3 5,=( )∧

where

l m, 1…2=( ) andl' 1 2l+=

m' 1 2m+=⎩⎨⎧

fM x( ) φl x( )

φ1 1 x L⁄–= and φ2 x L⁄=

Fes

fM x( )

Page 128: Seismic Assessment of Reinforced Concrete Frame Structures

104 Chapter 3

(3.20)

where the lower and upper integration limits are defined according to the control

section positions for the current load step.

However, in the present formulation, some of the integration limits may vary with the applied

moments. Recalling that element flexibility terms are defined as displacement derivatives with

respect to the moments, and that displacements result from the integration of deformations

between those varying limits, it becomes clear that contributions to the flexibility matrix may

be expected due to the internal moving sections.

In order to find out these additional contributions, one must look back at the definition of flex-

ibility terms in the present formulation. This will be restricted to the tangent flexibility

terms associated with an increment of end section moments after a previously equili-

brated moment distribution . For simplicity, no element applied loads are considered, as

their inclusion is straightforward.

The following derivation is based on an example where both yielding and cracking sections are

developed due to the applied increment of moments as shown in Figure 3.20-a). Since the

incremental form of the curvature distribution is essential for this derivation, both sets of yield-

ing and cracking sections, associated to the previous and the updated moment distributions, are

considered, i.e. and ; the corresponding flexibility distribu-

tions are included in Figure 3.20-b).

Curvature distributions shown in Figure 3.20-c) refer to the total ones, i.e. those necessary to

be provided by each section with the current flexibility in order to satisfy the equilibrium. For

clarity sake, the distributions of are also depicted, although their evaluation is not

actually needed along the whole element.

By definition, the tangent flexibility terms are given by

(3.21)

FlmM φl x( )fζ

M x( )φm x( ) xda1ζ

a2ζ

∫ζ 1=

Nzone

∑=

a1ζ a2ζ,( )

FlmM

∆MEm=1,2

M0 x( )

Y10 C1

0 C20 Y2

0, , ,( ) Y1 C1 C2 Y2, , ,( )

∆ϕζ x( )

FlmM

FlmM

∆MEm∂

∂∆θEl

M

=

Page 129: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 105

where the end section rotations , associated with the applied moments , result from

incremental curvature integration along the element as follows

(3.22)

Denoting , substituting Eq. (3.22) in Eq. (3.21) and recalling that

may vary with the applied moments, one obtains

(3.23)

The first term of this expression can be easily shown to be equivalent to Eq. (3.20) and the

zone division suggested by the distribution in Figure 3.20-b) would be enough for its

evaluation; instead, the second term accounts for the moving section contributions to the ele-

ment flexibility and is further detailed next.

According to Figure 3.20-a), ten integration zones are activated, the limits of which are

grouped as follows

It is apparent that only the abscissas related with may vary due to the current

increment of applied moments; therefore, denoting by the second term of Eq. (3.23), it

reduces to

(3.24)

Zone : 1 2 3 4 5 6 7 8 9 10

∆θEl

M ∆MEm

∆θEl

M φl x( )∆ϕ x( ) xd0

L

∫ φl x( )∆ϕζ x( ) xda1ζ

a2ζ

∫ζ 1=

Nzone

∑= =

glζ x( ) φl x( )∆ϕζ x( )=

a1ζ a2ζ,( )

FlmM

∆MEm∂

∂ glζ x( ) xda1ζ

a2ζ

∫ζ 1=

Nzone

∑ glζ a2ζ( )∆MEm

∂∂a2ζ glζ a1ζ( )

∆MEm∂∂a1ζ–

ζ 1=

Nzone

∑+=

fM x( )

ζ

a1ζ 0 xY1

0 xY1x

C10 xC1

xH xC2x

C20 xY2

xY2

0

a2ζ xY1

0 xY1x

C10 xC1

xH xC2x

C20 xY2

xY2

0 L

Y1 C1 C2 Y2, , ,( )

FlmM

FlmM φl xY1

( ) ∆ϕ2 xY1( ) ∆ϕ3 xY1

( )–[ ]∆MEm

∂xY1 φl xC1( ) ∆ϕ4 xC1

( ) ∆ϕ5 xC1( )–[ ]

∆MEm∂

∂xC1+=

φl xC2( ) ∆ϕ6 xC2

( ) ∆ϕ7 xC2( )–[ ]

∆MEm∂∂xC2 φl xY2

( ) ∆ϕ8 xY2( ) ∆ϕ9 xY2

( )–[ ]∆MEm

∂∂xY2+ +

Page 130: Seismic Assessment of Reinforced Concrete Frame Structures

106 Chapter 3

Note that the terms involving curvature increments are evaluated at both the left and

right sides of the same section; therefore, these increments are assigned the subscript corre-

sponding to the zone containing each side of the section.

Figure 3.20 Derivation of additional flexibility terms due to moving sections

Eq. (3.24) shows that, if no curvature discontinuities exist, as is the case for yielding sections,

the corresponding terms involving terms vanish and it simplifies to

∆ϕζ …( )

ζ

fM x( )

HE1 E2

M0 x( )

M x( )

fy1

- fy2

+

f01 f02

fp2

-fp1

-

C1C10

C20

Y10

Y20C2

Y1

Y2

C10 C2

0Y10 Y2

0

fy1

- fy2

+

f01 f02

fp2

-fp1

-

C1 C2Y1 Y2

fM0

x( )

∆ME1

∆ME2

C1

C2

Y1

Y2

C10

C20

Y10

Y20

ϕ0 x( )

ϕ x( )

ϕy1

-

ϕy2

+

∆ϕCc1

-

∆ϕCc2

+

∆ϕ x( )

xY1

0 xY1x

C10 xC1

0

L

xY2

0xY2x

C20xC2

1 2

7 8 9

3 45 6

10

xH

x

∆ϕζ x( )

ζ =

∆ϕζ x( )

1 2

7 8 9

3 4 5

6 10ζ =

a) Moment

b) Flexibility

c) Curvature

diagrams

diagrams

diagrams

∆ϕζ …( )

Page 131: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 107

(3.25)

in which the superscripts L and R have been introduced to identify the cracking section side

where the curvature increment is evaluated (following the convention stated in 3.5.3 for the left

and right sides).

From Eq. (3.25) it can be concluded that the additional terms arise due to the development

of cracking sections during the load step under analysis and to the “jump” in curvatures exist-

ing in the cracking section model. The first reason is related to the abscissa derivatives and

means that, if a cracking section is not developing, then no contribution is included in the flex-

ibility matrix. In turn, the second reason is associated with the difference of curvature incre-

ments , for , which reduces to the curvature “jump” referred in 3.5.2.

A similar reasoning can be applied to the case of full development of cracking sections and

as shown in Figure 3.19-a); the corresponding expression for the additional flexibility

terms becomes

(3.26)

where i refers to each element part and the unit factor takes the positive or negative sign for

and , respectively. Actually, Eq. (3.26) is still a particular case of the general

expression for the most complete case when , , and fully develop for para-

bolic moment distributions, which can be written in the following condensed form

(3.27)

where the counter has been introduced in accordance with Figure 3.6.

This allows to write the complete expression for the consistent tangent flexibility terms as

FlmM φl xC1

( ) ∆ϕC1

L ∆ϕC1

R–[ ]∆MEm

∂xC1 φl xC2( ) ∆ϕC2

L ∆ϕC2

R–[ ]∆MEm

∂xC2–=

FlmM

∆ϕCi

L ∆ϕCi

R–[ ] i 1 2,=( )

Ci

Ci+2

FlmM ξi φl xCi

( ) ∆ϕCi

L ∆ϕCi

R–[ ]∆MEm

∂xCi φl xCi 2+( ) ∆ϕCi 2+

L ∆ϕCi 2+

R–[ ]∆MEm

∂xCi 2++⎩ ⎭⎨ ⎬⎧ ⎫

i 1=

2

∑=

ξi

i 1= i 2=

Ci Ci+2 Ci+4 Ci+6

FlmM ξi φl xCi κ+

( ) ∆ϕCi κ+

L ∆ϕCi κ+

R–[ ]∆MEm

∂xCi κ+

κ 0 2 4 6, , ,=∑

⎩ ⎭⎨ ⎬⎧ ⎫

i 1=

2

∑=

κ

FlmM

Page 132: Seismic Assessment of Reinforced Concrete Frame Structures

108 Chapter 3

(3.28)

Closed form expressions can be obtained for the analytical calculation of each zone integrals

and the included derivatives are calculated upon definition of cracking section abscissas as

stated in 3.4.2 and in Appendix C. It must be noted that the additional flexibility terms have to

be obtained according to the applied loads (both end section and element loads).

According to Eq. (3.18), the terms are identified with , leading to the (6x6) matrix

, where the superscript M has been kept to remind the non-linear bending contribution to

the total matrix . Then, the transformation given by Eq. (3.8) is applied to and the

result is accumulated with the remaining contributions included in Appendix A, referring

already to the element local axis system.

3.6.4 Displacement residuals

The procedure for displacement residual computation as described in the general flexibility

formulation (see 2.4.4), typically requires the evaluation of section residual forces in order to

define residual deformations, whose integration along the element leads to the desired dis-

placement residuals.

Such procedure could also be used in the present formulation, but the moving character and the

simplified behaviour of internal control sections suggest an alternative methodology based on

the integration of total deformations rather than residual ones.

Consider again Figure 2.10 for the n-th Newton-Raphson iteration, aiming at target displace-

ments . For any internal iteration j, total deformations are given by

(3.29)

where for (j=1) and for (j>1). Total displacements are

then obtained as

FlmM φl x( )fζ

M x( )φm x( ) xda1ζ

a2ζ

∫ζ 1=

Nzone

∑ +=

ξi φl xCi κ+( ) ∆ϕCi κ+

L ∆ϕCi κ+

R–[ ]∆MEm

∂∂xCi κ+

κ 0 2 4 6, , ,=∑

⎩ ⎭⎨ ⎬⎧ ⎫

i 1=

2

FlmM Fl'm'

M

FesM

Fes FesM

un un-1 ∆un+=

etj x( ) ej-1 x( ) ∆ej x( ) ∆er

j x( )+ +=

ej-1 x( ) en-1 x( )= ej-1 x( ) etj-1 x( )=

Page 133: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 109

(3.30)

and, it is intuitive, that residual displacements can simply be obtained by the difference of

to the target ones . Actually, the following alternative expression

(3.31)

can be proved to be equivalent to

(3.32)

therefore, leading to another way of computing displacement residuals, provided the total

deformations at the section level are known.

Should the global N-R scheme be based on incremental iterative corrections of displacements

relative to the step beginning as suggested in 2.4.4.4, the same procedure is still valid if refer-

ences to and to are replaced by and , respectively. Accordingly, total defor-

mations are also computed by accumulation over (for j=1),

i.e., over a deformed shape corresponding to a duly converged equilibrium configuration.

Since only one bending direction is assumed to behave non-linearly and element applied loads

are only considered acting on the corresponding plane, displacement residuals are likely to

develop only for components affecting that direction. According to the notation of 3.2, the lin-

early independent vectors , and of Eq. (3.9) refer to components behaving linear

elastically and the corresponding residual displacements are null; therefore, the result of defor-

mation integration leads to the input displacements, which means that such integration is use-

less for those terms. The remaining vector includes the non-linear displacement

components, so it has to be obtained by integration of the relevant deformations (see 3.6.7),

namely the shear distortion and the bending curvature in the xz plane. The former contribution

is straightforward due to its linear behaviour, whereas the latter is more demanding because the

curvature distribution has to be adequately set up according to the non-linear behaviour.

utj bT x( ) et

j x( )⋅ xd0

L

∫=

∆urj

utj un

∆urj ut

j un– bT x( ) etj x( )⋅ xd

0

L

∫ un-1 ∆un+( )–= =

∆urj bT x( ) ∆er

j x( )⋅ xd0

L

∫=

un-1 ∆un u0 ∆u0n

etj x( ) ej-1 x( ) e0 x( ) ek-1 x( )= =

uux u

θx uθz

uθy

Page 134: Seismic Assessment of Reinforced Concrete Frame Structures

110 Chapter 3

This alternative methodology for residual computation has been adopted herein for its ade-

quacy in the present formulation context. The following reasons sustain this option:

• The model characteristics associated to internal moving sections are such that total curva-

tures can be readily known for given applied moments; moreover, this is valid for all

uncracked and cracked zones, which usually form a major part of the element.

• If residual curvature distribution were to be obtained, then yielding and cracking

sections of both the current and the previous iterations would have to be controlled, as

shown in Figure 3.20; note that in this figure reference is actually made to the total curva-

ture increment , which includes the predictor and the residual/corrector incre-

ments, and the whole set of zones would be required to perform integration.

• By contrast, if only total curvature distribution is of interest, there is no need for controlling

the previous iteration moving sections, due to the uniform, and readily known, flexibility

distribution along uncracked and cracked zones. This leads to an important economy from

the computation time standpoint, since significantly less internal moving sections are actu-

ally controlled.

Therefore, it is clear that for most of the element zones, residual computation via total curva-

tures is readily achieved. However, to cope with this strategy, special care has to be taken with

the yielded zones, since the total curvature distribution definition is not straightforward when

unloading or reloading situations occur in such zones. This aspect is specifically detailed in

3.6.6 where a special procedure is proposed to improve consistency between distributions of

applied moments and total curvatures, for the assumed flexibility distribution.

Additionally, the global N-R scheme adopted in this work is actually based on incremental iter-

ative corrections relative to the step beginning, which is compatible with the fact that the

cracking and yielding section development is irreversible between equilibrated steps, though

not necessarily between iterations. In other words, the section movement is irreversible with

respect to their position at the step beginning, but, between consecutive internal iterations j-1

and j of a given external N-R iteration n, their positions can change regardless of keeping or

reversing the direction of movement. Thus, once cracking and yielding sections are found for a

given internal iteration j, their irreversible movement conditions must be checked with respect

to the step beginning and, according to Figures 3.6 and 3.9, they can be expressed in terms of

local abscissas by the following relations:

∆ϕr x( )( )

∆ϕ x( )( )

Page 135: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 111

where the superscript 0 stands for the step beginning and the cases of sections and

have been included for completeness, although they are not considered in this study.

3.6.5 Element applied loads

In the present work, element applied loads have been considered according to the Spacone pro-

posal (Spacone (1994)) as described by Eqs. (2.28) to (2.34) in 2.4.4.3.

Since these loads are assumed fully applied at the first load step, thus for elastic behaviour, that

proposal becomes quite suitable and efficient from the computational standpoint. In fact, the

element fixed-end forces due to element applied loads, either distributed p or concentrated

P at section H, can be readily obtained by Eq. (2.32) with the elastic flexibility distribution

, as defined in Appendix A by Eq. (A.4), and with the corresponding stiffness matrix .

Particularly, the integral of Eq. (2.32) gives the elastic displacements due to total element

applied loads in the reduced space, which can be calculated by the closed form expressions

included in Appendix A (see A.3), accounting for the type of load and the distribution of sec-

tion properties along the element.

Thereby, Eqs. (2.32) and (2.33), for the total element loads (p instead of ), can be re-written

respectively as

(3.33)

and

(3.34)

which hold only for the first internal iteration (j=1) of every N-R iteration n, according to the

adopted global N-R scheme referring to the step beginning.

For subsequent internal iterations (j>1) the contribution of is removed from Eq. (3.34), but

the distribution is kept explicitly included in order to correctly evaluate forces in inter-

sYi

j sYi

0≥ sCi

j sCi

0≥ sCi 2+

j sCi 2+

0≤ sCi 4+

j sCi 4+

0≥ sCi 6+

j sCi 6+

0≤

Ci 4+ Ci 6+

Qp

f0 x( ) K0

up

∆p

Qp K0– up⋅=

∆S x( )j=1 b x( ) K0 ∆un up–( )⋅[ ]⋅ Sp x( )+=

up

Sp x( )

Page 136: Seismic Assessment of Reinforced Concrete Frame Structures

112 Chapter 3

nal sections. Expressions of (and of in the local axis system) are included in

Appendix C for both the distributed load p and the concentrated force P at the span section H.

It is reminded that, once the restoring element forces are obtained in the reduced space, the

transformation to the element space with rigid body modes is performed by Eq. (2.11), where

the contribution of element applied loads must be included in every N-R iteration n of the

first step; again, expressions of are given in Appendix C for both types of element loads.

3.6.6 Behaviour of plastic end zones

3.6.6.1 General

The behaviour of yielded zones (herein also designated as plastic zones) needs to be carefully

analysed, since it is effectively controlled only at their boundary sections, i.e., end and yielding

sections, and nothing is known concerning the intermediate sections for most loading stages.

For the assessment of curvature distribution in plastic zones, the following loading case dis-

tinction is important:

i) post-yielding loading responsible for monotonic development of the plastic zone

ii) unloading, reloading or post-yielding loading for which the plastic zone has not been

continuously developing.

The loading case i) is easily handled since all the intermediate sections have their state points

lying on the model post-yielding branch. This is illustrated in Figure 3.21 for the simple case

of a cantilever with a uniformly distributed load and a vertical force applied at the tip, in which

the yielding moment has been exceeded. Both moment, flexibility and total curvature distribu-

tions are included, as well as the basic model curve (assumed uniform along the element).

For a linear moment diagram, the total curvature distribution in the plastic zone (labelled by 1

in Figure 3.21) is readily known once the state determination is performed for its boundary

sections ( and ). In turn, for the case of the parabolic moment diagram (see Figure 3.21),

the coefficient of the 2nd order term of the curvature distribution is easily shown to

depend only on the distributed load and on the post-yielding stiffness . This coefficient

remains constant as long as the distributed load does not change, while the other coefficients

( and ) can be obtained through the total curvature values at the end and yielding sec-

Sp x( ) Spsx( )

Qn

Qpf

Qpf

E1 Y1

aϕ1( )

p fp

bϕ1cϕ1

Page 137: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 113

tions (respectively, and ) in order to completely define the distribution .

Figure 3.21 includes also the curvature distributions and for the remaining zones

(cracked and uncracked ones, respectively) as well as their second order polynomial coeffi-

cients; once these coefficients are known, along with curvature values at each zone boundary

sections, the curvature distribution can be completely defined.

Figure 3.21 Monotonic development of plastic zone

It is noticed that, even for subsequent loading stages as those of ii), plastic zones have to pass

through the above shown curvature diagram because yielding development is only assumed to

start from the end sections. Therefore, for any loading process involving distributed load at the

beginning, the very first coefficient is always as indicated in Figure 3.21.

The loading cases of ii) are such that total curvature distribution in plastic zones cannot be

ϕE ϕy ϕ1 x( )

ϕ2 x( ) ϕ3 x( )

Mc

C1

E1 E2Y1

My

fy f0

fp

M x( ) aMx2 bMx cM+ +=

f x( )

MEM x( )

ϕ

M

oϕcc

ϕc

fpE

ϕy

ϕccϕc

ϕy

ϕ1 x( )

ϕ1 x( ) ϕy fp M x( ) My–[ ]+=

C1Y1

ϕ2 x( ) fyM x( )=

ϕ3 x( ) f0M x( )=

ϕ1 x( )

ϕE

aϕ1fpaM=

ϕζ x( ) aϕζx2 bϕζ

x cϕζ+ +=

aϕ2fyaM=

aϕ3f0aM=

p

F

ζ=1 2 3

ζ 1 2 3, ,=( )

a) Moment

b) Flexibility

d) Curvature

diagramc) Moment-Curvature

aMp2---=

aϕ1

Page 138: Seismic Assessment of Reinforced Concrete Frame Structures

114 Chapter 3

explicitly derived from the model. The following general expression may be considered

(3.35)

which can be obtained from the expressions included in Figure 2.10, after particularizing for

curvatures. The predictor term, based on the previous flexibility and the increment of

applied moments , is added with the corrective term due to unbalanced moments

along with the updated flexibility . As previously stated, flexibility is assumed

linearly distributed, but information is still lacking concerning , since the state points

in the model (i.e., resisting moments ) for intermediate sections are unknown and, gen-

erally, cannot be inferred from those of end and yielding sections.

Actually, for certain simple loading cases (as the first unloading from post-yielding branch, for

instance), a relation could be derived between the assumed and the consistent ,

within the model rules; however, once loading becomes more complex, such relation turns out

to be impossible to obtain. Another possibility could be to assume a distribution law for

, which, however, should be consistent with the adopted for all loading stages.

Since this is practically impossible to be achieved, such option could not ensure solution objec-

tivity and algorithm convergence.

In the present study this problem is overcome by the use of an event-to-event type procedure

inside the iterative process for the element state determination. The event-to-event technique

(Simons and Powell (1982), Porter and Powell (1971)) is suitable for non-linear problems

behaving linearly after and before specific events. By event it is meant a stiffness variation

which can be predicted according to the model rules and the current state. The basic idea is to

closely follow the model path at all loading steps, by adapting them to the event sequence and

updating the stiffness and state accordingly, in order to eliminate unbalanced forces. This tech-

nique is further detailed in Appendix D with the help of a very simple illustrative example; its

application to the present element state determination is explained later in this section.

In the present context, the event-to-event scheme is used only for the control of sections gov-

erning the plastic zone behaviour. The main purpose is to assure that, a sub-increment being

applied at each iteration, the flexibility distribution inside plastic zones either remains

unchanged or modifies to but with negligible unbalanced moments ; in both

ϕj x( ) ϕj-1 x( ) fj-1 x( )∆Mj x( ) fj x( )∆Muj x( )+ +=

fj-1 x( )

∆Mj x( )

∆Muj x( ) fj x( )

∆Muj x( )

Mrj x( )

fj x( ) Mrj x( )

Mrj x( ) fj x( )

fj-1 x( )

fj x( ) ∆Muj x( )

Page 139: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 115

cases, the following approximation holds

(3.36)

In fact, behind this approach there lies the assumption that, if flexibility remains unchanged at

the plastic zone boundary sections, the same happens for all the yielded sections and, therefore,

residuals become null. Obviously, this is enforced by the adopted linear flexibility approach

and, for complex loading situations, the realism of such option may be questionable from a

local section standpoint. Nevertheless, it should be recalled that, rather than a local section

detailed behaviour, a global element response is to be assessed; in such context, this approach

is believed to be acceptable, although it is recognized that improvements may be needed in

future developments.

3.6.6.2 Plastic zone splitting

It is important to realize that, for the flexibility distribution to remain unchanged in the

plastic zones, it is necessary that neither the boundary section flexibilities nor the zone length

change during a given sub-increment. The first requirement is accounted for with the event-to-

event scheme, whereas the second one is achieved by dividing each plastic zone into two parts

as shown in Figure 3.22. One fixed part between the end section and the yielding one

of the previous step (duly equilibrated and converged) and a variable length part,

between and the yielding section of the present iteration.

Figure 3.22 Plastic zone splitting in fixed and variable length parts

With this plastic zone division it follows that:

• variable length parts are straightforwardly handled as explained above for the monotonic

development of plastic zones (see Figure 3.21);

• the event-to-event procedure is applied to control the fixed parts, specifically at the end sec-

ϕj x( ) ϕj-1 x( ) fj-1 x( )∆Mj x( )+≈

f x( )

Ei=1,2( )

Yi=1,20( )

Yi=1,20 Yi=1,2( )

HE1 E2

M0 x( )

M x( )

Y10

Y20

Y1

Y2

∆ME1

∆ME2

Fixed Variable Length

Page 140: Seismic Assessment of Reinforced Concrete Frame Structures

116 Chapter 3

tions and the left sides of .

Depending on the state of boundary sections, flexibility distributions in fixed plastic zones can

develop in several possible ways as shown in Figures 3.23 and 3.24, for a given load step.

Dashed line diagrams represent the flexibility corresponding to the previous equilibrated

step, whereas solid line diagrams refer to the updated flexibility ; for simplicity it is

assumed that at most one event can occur at those boundary sections during the load step. Pos-

sible positions of those section state points in the moment-curvature diagram are also included

and small arrows indicate the post-event evolution.

Figure 3.23 Flexibility distributions in plastic zones with no further yielding development

Figure 3.23 shows some cases with no yielding development, which are briefly described next:

Yi=1,20

f0 x( )

f x( )

Y0 Y≡E

M

o

E

fy+

fp+

fu+

ϕ

YL0

Events in: E & Y0

Y0 Y≡E

o E

fy+

fr1

-

fu+

Y0

Events in: E & Y0

fy-

fr1

-fy-

Y0 Y≡E

E

fr1

-

fp-

Y0Event in: E

fy-

fp-

fy-

Y0 Y≡E

E

fr1

-

Y0

Event in: E

fy-

fy-

fr2

-

fr2

-

Y0 Y≡E

E

fy+

fp+

YL0

Event in: Y0

Y0 Y≡E

E

fy+

fp+

YL0

Events in: E & Y0

fu/r+

fu/r+

fy+ fu

+

fp+

f x( )f0 x( )

b)a)

d)c)

f)e)

Page 141: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 117

• Figure 3.23-a) refers to unloading of both E and sections from the post-yielding branch,

which means events taking place at both sections. Therefore, a very small (“infinitesimal”)

sub-increment is first applied with the flexibility, just to enforce the changing to ;

the remaining displacement sub-increment is then applied with the already updated flexibil-

ity distribution. No residuals are expected as long as no more events take place, but if this

should happen, as a result of flexibility modifications, a new subdivision would be

enforced.

• Figure 3.23-b) shows the case of events taking place at both boundary sections due to

changing of moment sign. The first sub-increment is applied with until zero moments

are reached, after which flexibility can be switched to , in order to apply the remaining

sub-increment.

• Figures 3.23-c) and 3.23-d) represent reloading cases where the end section is likely to

reach again the post-yielding branch or to change stiffness due to pinching effects, whereas

remains in the same branch.

• Figures 3.23-e) and 3.23-f) refer to situations of opposite loading direction in boundary sec-

tions; while unloads, the end section either remains on the post-yielding branch (case e))

or reaches it from a reloading/unloading situation (case f)).

Figure 3.24 Flexibility distributions in plastic zones with further yielding development

Y0

f0 x( ) f x( )

f0 x( )

f x( )

Y0

Y0

Y0E

fy+

fp+

YY0E

M

o

E

fy+

fp+

ϕ

YL0

No Events

EY0Event in: Y0

EYL0

Event in: E

Y0E

E

fy+

fp+

YL0

Event in: E

fu/r+

fu/r+

Y

fu+

fu+

Y0E

fy+

fp+

Y

fu/r+

Y

fp+

fy+

b)a)

d)c)

Page 142: Seismic Assessment of Reinforced Concrete Frame Structures

118 Chapter 3

Some yielding development situations are illustrated in Figure 3.24, namely:

• The case a), which is the same of Figure 3.21, but it is irrelevant for events detection.

• Unloading at the end section from the post-yielding branch, while still loads along it, as

in the case b); therefore, an event at section E enforces the sub-incrementation process with

a procedure similar to the case of Figure 3.23-a).

• The situation c), where the end section remains in a reloading/unloading phase, while

changes to the post-yielding behaviour; until this event occurs in , a first sub-increment is

applied, after which the flexibility is updated.

• The case d), in which an end section event occurs when it reaches again the post-yielding

branch after a reloading/unloading situation; the left side of keeps on loading along the

same path.

3.6.6.3 Event-to-event scheme in the element iterative process

As previously mentioned, the event-to-event scheme is used in the present work, to control the

boundary sections of fixed plastic zones ( and ), because the assumed step-wise lin-

ear behaviour is suitable to be controlled by that technique. A similar scheme has been used by

Filippou and Issa (1988) but with the purpose of controlling the plastic zone development.

It is worth stressing the contrast between the event-to-event scheme used at the global structure

level (as described in Appendix D) and its application at the element level. In the first case, any

event enforces a subdivision of the global load increment and the whole incremental/iterative

process may easily become inefficient when strongly non-linear behaviour is encountered. In

turn, the element event-to-event scheme includes a very limited number of sections generating

events (four, at most) and it is activated only inside the internal element iterative process, thus

without directly affecting the global structure N-R process.

In other words, the advantage of using the event-to-event scheme herein is twofold: a) it pro-

vides a suitable means of achieving the state determination of the fixed plastic zones (within

the assumptions stated in 3.6.6.1) and b), being used at the element level, it is not likely to

affect significantly the global N-R scheme.

The sequence for the event-to-event application is schematically shown in Figure 3.25. This

scheme becomes more efficient if there is at most one event per section; therefore, prior to the

sub-incrementation for reaching an event, the auxiliary increment is made equal to the

Y0

Y0

Y0

Y0

Ei=1,2 Yi=1,20

∆ua( )

Page 143: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 119

total increment and subdivided into two parts ( and ) such that no more than one

event is found at each section upon application of (see step 1.1/ of Figure 3.25).

For the application of , events are searched among the boundary sections of fixed plastic

zones. Following a procedure similar to that explained in Appendix D, increment reduction

factors are set up for those sections, and the event (V), if existing, is associated to the minimum

factor. Thus, in case of event, is first considered as the working increment and is

split into two parts:

Figure 3.25 Application of the event-to-event scheme to the element state determination

• the first sub-increment is applied with the previous flexibility, up to the turning

point, but slightly exceeding it; this enforces the flexibility variation with numerically neg-

ligible unbalanced moments at the boundary sections of plastic zones, but it does not ensure

∆u( ) ∆ub ∆ur

∆ub

∆ub

∆ub ∆uw( )

V

VTarget

Preliminary Subdivision

Events Detection in all sections

Sub-incrementation

State determination for

Set up:

Set:

Residuals can appear

∆u Total Displacement Increment=

∆ub∆ur

u0 ufub

u0 ∆u1

∆u2δu1∆u1

δ– u1

∆uw ∆u2 δu1–=

δu1( )

∆uw ∆ub=

Check Convergence for

(Max. 1 event per section)

∆ua ∆u=

∆ua

∆uw

∆u2

1/

1.1/

2/

2.1/

2.3/

2.2/

V = event closest to the beginning

∆u1

Update flexibility distribution

∆uw ∆u2 δu1–=

∆uw - NON Converged => Restart 2.1/

- Converged -> ∆ur 0=( ) STOP⇒

∆ur 0≠( ) ∆ua ∆ur=( )⇒Restart 1.1/

If

If

u1

∆u1( )

Page 144: Seismic Assessment of Reinforced Concrete Frame Structures

120 Chapter 3

the increment of deformations along the element to be compatible with ; indeed, modi-

fications on the stiffness and positions of the remaining internal sections (yielding and

cracking ones) may lead to element displacements different from , the difference being

designated by residual displacements ;

• the second sub-increment is applied with the updated flexibility, but corrected with

residuals resulting from the first sub-increment.

The application of is followed by the setting up of moving sections and the state determi-

nation at every section; the flexibility distribution is updated and displacement residuals

are computed. In order to respect the target displacements (see steps 2.1/ and 2.2/ of Figure

3.25), the residual correction is then applied together with , if still existing.

Thus, a new working increment is given by , where only residuals may per-

sist in case the sub-incrementation has not been done. This second and the following phases for

application of must be performed again from steps 2.1/ in order to enforce sub-incremen-

tation for further events likely to develop; the process stops when convergence is reached as

stated in 3.6.8.

According to 3.6.4, displacement residuals are calculated via integration of total curvatures;

thereby, after application of , total displacements are obtained and

the new working increment can be readily calculated from the alternative expression

, where actually stands for the updated displacements at the end of

each iteration.

After completion of , both element and section states are duly updated and, if there is still

any remaining increment part to be applied, the process restarts from step 1.1/.

It is worth mentioning that, increment subdivisions are performed directly on the applied dis-

placements, instead of the corresponding moments because these are dependent on the flexibil-

ity matrix used in the state determination.

3.6.6.4 Evolution of curvatures in fixed plastic zones

Application of Eq. (3.36), for total curvature diagram definition, is actually performed only in

the fixed plastic zones. It is apparent that accumulation of deformations is taking place from

∆u1

∆u1

δu1

∆u2( )

δu1( )

∆u1

δu1( )

δ– u1( ) ∆u2

∆uw ∆u2 δu1–=

∆uw

∆u1 u1 u0 ∆u1 δu1+ +=( )

∆uw ∆ub u1 u0–( )–= u1

∆ub

∆ur

Page 145: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 121

one iteration to another, which must be reflected in the coefficients of the curvature equation.

If element applied loads are assumed to take place only at the first load step, as referred to in

3.1, then further step loading or iterative corrections consist only of end section applied loads,

leading to linear functions for . With the linear flexibility distribution approach, Eq.

(3.36) implies curvature increments to have (at most) a 2nd order term, whose coefficient is

readily obtained from the expressions of and . This is shown in Figure 3.26 for the

same example of Figure 3.21, where unloading is now taking place due to a tip force applied in

the reverse direction.

Figure 3.26 Total curvature evolution for non-monotonic loading

According to Figure 3.26, the 2nd order coefficient of the parabolic increment of curvatures

is added to the value corresponding to the previous iteration (or step), in order to

obtain the coefficient for the updated total curvature expression; the remaining coefficients

∆M x( )

f x( ) ∆M x( )

E1 E2

fy f0

∆M x( ) b∆x c∆+=

f x( )

ME ME

ϕ

M

oϕcc

ϕc

fp

E

ϕy

ϕccϕc

ϕE

ϕE ϕE0 fu ME ME

0–( )+=

ϕ2 x( ) fyM x( )=

ϕ3 x( ) f0M x( )=

ϕ1 x( ) ϕ10 x( ) f1 x( )∆M x( )+=

ϕE

p

aϕ1aϕ1

0 bfb∆+=

ϕζ x( ) aϕζx2 bϕζ

x cϕζ+ +=

aϕ2fyaM=

aϕ3f0aM=

F

M0 x( )

f1 x( ) bfx cf+=

M x( )

ϕy

Y10 C1

0

fu

Y10 C1

0

ME0

ϕE0

fuME

0

ϕE0

ζ=1 2 3

ζ 1 2 3, ,=( )

a) Moment

b) Flexibility

d) Curvature

diagramc) Moment-Curvature

aMp2---=

bfb∆( ) aϕ1

0

aϕ1

Page 146: Seismic Assessment of Reinforced Concrete Frame Structures

122 Chapter 3

can be derived from the curvature values at sections and .

Although the case of element applied loads in other steps (after the first one) is not considered

in the present study, it is clear from Figure 3.26 that it would lead to a 3rd order polynomial for

the curvature diagram. In such a case, the 3rd and 2nd order coefficients for the curvature incre-

ment due to an element load increment , would be computed only in terms of coeffi-

cients and of and then accumulated with those from previous steps. For the remaining

coefficients, the procedure referred to in the previous paragraph applies.

It is apparent that for no inclusion of distributed load p, the second order coefficients are

restricted to the plastic zone, having only the contribution of generated once unloading

has taken place.

3.6.7 Integration of deformations

Once the curvature distribution is adequately updated, the integration is performed over the

distributions of deformations relevant for the non-linear displacements included in . How-

ever, since both moment and curvature distributions have been referred to the local section axis

system, the vector is actually obtained first.

Looking back at Eqs. (3.9) and (3.10) and following a procedure similar to that of 3.6.3, the

two non-zero components of (i.e., the 3rd and the 5th ones ) can be given by

(3.37)

where the superscripts V and M stand for the shear and the bending contributions, respectively.

According to Eq. (3.22), the terms are given by

(3.38)

for which, the integrals of each zone can be analytically evaluated by closed form expressions.

Similarly, the shear related terms are defined by

E1 Y10

∆p( ) f x( )

∆p

aϕζ

bfb∆

uθy

uesθy

uesθy

θEl'θEl

θEl

V θEl

M+= = l 1 2,= l' 1 2l+=→( )

θEl

M

θEl

M φl x( )ϕ x( ) xd0

L

∫ φl x( )ϕζ x( ) xda1ζ

a2ζ

∫ζ 1=

Nzone

∑= =

θEl

V

Page 147: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 123

(3.39)

where is the shear distortion distribution and are the shear force shape functions,

which can be drawn from Eq. (3.6) and are given by

(3.40)

Since the shear behaviour is linear, holds, where remains constant

during the loading process and is assumed as given by Eqs. (A.5) and (A.6) in Appendix A.

The distribution may include element applied loads and is expressed by Eqs. (C.7) and

(C.8) in Appendix C. Therefore, Eq. (3.39) becomes

(3.41)

where the (-) sign holds for and the (+) for , and .

According to the notation of Appendix A, the above integrals yield

(3.42)

which, for the case of uniform properties along the element, simplifies to .

After computation of as above explained, Eq. (3.3) is applied to transform it to ,

referred to the element axis system x,y,z. Then, the result is superimposed to the remaining

components of Eqs. (3.9) and (3.10) to obtain the complete displacement vector . However,

note that such operation is worthless from the algorithm standpoint, since residuals and, conse-

quently, convergence have to be checked only for the non-linear components, i.e., directly for

. On the other hand, for the element output forces, the superposition of the non-linear com-

ponents with the linear ones is relevant and has to be performed.

θEl

V φlV x( )β x( ) xd

0

L

∫=

β x( ) φlV x( )

φ1V 1 L⁄–= and φ2

V 1 L⁄=

β x( ) fV x( )V x( )= fV x( )

V x( )

θEl

V 1L---± 1

GAz( )1----------------- VE p L

2--- x–⎝ ⎠⎛ ⎞– P 1 h–( )– xd

0

hL∫

1GAz( )2

----------------- VE p L2--- x–⎝ ⎠⎛ ⎞– Ph– xd

hL

hL

+⎩

=

l 1= l 2= VE ME2ME1

–( ) L⁄=

θEl

V hGAz( )1

----------------- 1 h–( )GAz( )2

-----------------+ VE1

GAz( )1----------------- 1

GAz( )2-----------------– h2 h–( ) pL

2------ P+⎝ ⎠⎛ ⎞+

⎩ ⎭⎨ ⎬⎧ ⎫

±=

θEl

V 1GAz----------VE±=

uesθy u

θy

u

uθy

Page 148: Seismic Assessment of Reinforced Concrete Frame Structures

124 Chapter 3

3.6.8 Convergence criteria for the element iterative process

The internal element iterative scheme aims at calculating restoring forces for a given increment

of applied displacements, by means of progressive elimination of displacement residuals. Con-

vergence is reached when these residuals are smaller than a pre-defined measure of numeri-

cally negligible displacements. Convergence is checked only in terms of rotations in the non-

linear bending plane; the corresponding tolerance for rotation residuals is denoted here by

which is controlled by a user-supplied relative tolerance (e.g. 10-4, 10-6, ...).

Therefore, convergence is considered to be reached when the following condition is satisfied

(3.43)

where stand for residuals of element end section rotations.

The value of has to be chosen according to several distinct situations likely to develop in

the element. Thereby, in the present work it is given by the following condition

(3.44)

where:

• is the maximum absolute value of rotation increments applied to the element in the

current step;

• is the maximum elastic rotation (absolute value) corresponding to the element end sec-

tion moments at the step beginning;

• is the maximum elastic rotation (absolute value) associated with anti-symmetric bending

produced by end section moments equal to the average of absolute values of cracking

moments in the element and

• is a value close to (but slightly higher than) the numerical precision of the computer.

Comparison of against is the criterion often used; however, for very small (or even

null increments) this condition is not applicable, for which the displacement measure

related with the installed stress state may be more appropriate. In the case of initial step, where

only element loads p (or P at section H) are considered, the structure is unstressed (thus,

) and no displacement increments are considered (i.e., ); therefore, the con-

θtoler ∅

θrmax max ∆θr1

∆θr2( , ) θtoler<=

∆θri=1,2

θtoler

θtoler max ∅∆θa( ) ∅θe0( ) ∅θc( ) Precis, ,{ , }=

∆θa

θe0

θc

Mc

Mc Mc1

+ Mc1

- Mc2

+ Mc2

-+ + +( ) 4⁄=( )

Precis

θrmax ∅∆θa

θe0

θe0 0= ∆θa 0=

Page 149: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 125

vergence check against and fails and another criterion related to the internal

resistance of the element is adopted. The rotation has been chosen as a kinematic measure

of the cracking limit and the verification is done for . Finally, none of the previous rota-

tion tolerances can be lower than the computer precision and the is imposed as the

lower bound of tolerance.

3.6.9 Convergence problems

3.6.9.1 Difficult or no- convergence situations

The particular features of the model to overcome the cracking transition may lead to difficult

convergence situations that are likely to develop depending on the moment distribution along

the element.

Simple cases, as for instance a cantilever beam loaded by a tip force or a beam under anti-sym-

metric bending, do not present problems related with cracking transition, provided the addi-

tional flexibility terms are duly incorporated in the flexibility matrix as detailed in 3.6.3.

Figure 3.27 shows such the simple case of a cantilever beam subjected to linear bending

moment distribution , behaving as uncracked for the step 0. By imposing a displacement

increment with the previous stiffness (the elastic one), the moment distribution for the first

internal iteration (j=1) becomes (see Figure 3.27-a)), inducing cracking initiation.

The corresponding flexibility and total curvature distributions (see Figures 3.27-c) and 3.27-

d)) are dictated by the section model (shown in Figure 3.27-b)) and are denoted, respectively,

by and . Total displacements can be obtained upon integration of

and the corresponding displacement residuals can be evaluated. The updated flexibility

matrix arises from integration of , corrected with the additional flexibility terms

due to cracking section movement.

The second internal iteration (j=2) is then performed by imposing the correction of moments

leading to and obtained from and the previous stiffness matrix , which

accounts for the section stiffness variation ( to ) and for the curvature “jump” in the

cracked zone developed. The presence of the additional flexibility terms ensures that the ele-

ment stiffness drops sufficiently to keep the end section state point above the crack-

ing plateau (see Figure 3.27-b)); thereby, cracking is still detected in the element and the

∅∆θa ∅θe0

θc

∅θc

Precis

M0 x( )

∆u

Mj=1 x( )

fj=1 x( ) ϕj=1 x( ) uj=1 ϕj=1 x( )

∆urj=1

Fj=1 fj=1 x( )

FlmM

Mj=2 x( ) ∆urj=1– Kj=1

f0 fy

Kj=1 Ej=2

Page 150: Seismic Assessment of Reinforced Concrete Frame Structures

126 Chapter 3

successive iterations tend to adjust the cracked zone until the total displacements obtained

from integration match the desired ones (i.e., ) within a pre-defined toler-

ance.

Figure 3.27 Cracking transition and the role of additional flexibility terms

However, if these additional terms were not included, the flexibility matrix would contain only

the influence of section flexibility variation from to , which could also occur if the M-ϕ

diagram followed the dashed line (parallel to the cracked branch) after point C. In these condi-

tions, the element flexibility matrix would not “recognize” the curvature “jump” along the

cracked zone, which means an inconsistency between that matrix and the developed curva-

tures. Consequently, a higher element stiffness would arise and, depending of the magnitude of

the curvature “jump” , the enforced correction to could lead to a state point

below the cracking plateau (see Figure 3.27-b)). The elastic behaviour would be

reached again, the element stiffness would turn back to the initial state and the next iteration

(j=3) would develop as the first one. A closed loop would take place, switching between two

opposite states (below and above the cracking plateau) without the possibility of reaching con-

vergence.

uconverg u0 ∆u+≅

Mc

C1

E1 E2

fy f0

Mj=1 x( )

fj=1 x( )

ME

ϕ

M

oϕcc

ϕc

C

ϕccϕc

ϕE

C1

ϕj=1 x( )

∆urj=1 ∆uj=1 ∆u– uj=1 u0 ∆u+( )–= =

a) Moment

c) Flexibility

d) Curvature

b) Moment-Curvature diagram

M0 x( )

Mj=2 x( )Ej=1

Ej=2

Ej=2( )'

fy

f0

∆MEj=2 ∆ME

j=2( )'

C1

ϕ0 x( )

∆uj=1

∆u ∆MEj=1⇒

f0 fy

∆ϕccMj=2 x( )

Ej=2( )'

Page 151: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 127

Therefore, the inclusion of additional flexibility terms ensures the consistency between the ele-

ment flexibility and curvatures and avoids the above referred closed loop problem.

On the other hand, it must be realized that such terms become less relevant for the element

flexibility as the cracking sections move farther away from the end ones. Taking the simple

example under study, and considering the stage where only the section is developing, the

additional term expression (see Eq. (3.25)) reduces to

(3.45)

where the term coincides with the cracking curvature “jump” .

Particularly for the terms , which are the most affected by the cracking development next

to the end section , Eq. (3.45) can be written as

(3.46)

showing that increasing values of ( moving apart from ) tend to reduce . In turn,

the lower or higher variation of with the applied moments is expressed by the derivatives

and depends mainly of the moment diagram slope in that zone.

This means that, for an “almost flat” moment distribution in element zones crossing the crack-

ing plateau, sudden and significant development of a cracked zone may occur, whereby the

additional terms can be of less relative weight in the element flexibility matrix. Therefore,

a situation of closed loop as above referred can be triggered off by lack of flexibility for the

applied moment increment (or iterative correction).

Two typical situations of this type are shown in Figure 3.28. In the first one (see Figure 3.28-

a)) an initially uncracked element is forced to develop a large cracked zone due to small end

section increments leading to a very flat diagram; such situation requires a pronounced stiff-

ness decrease in order to compensate for the large residuals developed in the cracked zone.

Another case is shown in Figure 3.28-b) where an existing slightly cracked element part may

C1

FlmM φl xC1

( ) ∆ϕC1

L ∆ϕC1

R–[ ]∆MEm

∂∂xC1=

∆ϕC1

L ∆ϕC1

R–[ ] ∆ϕcc

F1mM

E1

F1mM 1

xC1

L-------–⎝ ⎠

⎛ ⎞∆MEm

∂xC1 ∆ϕcc=

xC1C1 E1 F1m

M

xC1

∂xC1∂∆MEm⁄

F1mM

Page 152: Seismic Assessment of Reinforced Concrete Frame Structures

128 Chapter 3

become fully-cracked due to small variations of end section moments generating an almost flat

diagram in that element part ( and are supposed of having been generated by previous

moment distributions).

Note that such problems are not caused by inconsistencies in the formulation and a very strin-

gent subdivision of the applied increment would lead to a converged solution, but at a quite

high computational cost.

Similar convergence difficulties are reported in the literature (Criesfield (1982)) often related

with reinforced concrete cracking problems, which are solved by adopting line search tech-

niques. This has been the strategy adopted also in the present work to overcome the cracking

plateau transition. As the event-to-event scheme referred in 3.6.6, a line search procedure can

be very efficient in the present context because it is restricted to the element level rather than

involving the state determination for the entire structure.

Figure 3.28 Typical cases generating convergence problems

3.6.9.2 Line search scheme for element iterations

The main and basic features of line search schemes are briefly recalled in the next paragraphs,

devoting particular attention to the adaptation for the present element iterative algorithm and to

the criteria for the selection of the increment scaling factor.

Line search schemes are based on the recognition that the increment, or iterative correction, of

C1 C3

Mc

C1

E1 E2H

M x( )

a)

C1

M0 x( )

H

H

E1 E2

Mc

E1 E2H

M x( )

b)

M0 x( )

H

H

E1 E2C1 C3

for M0 x( )

for M x( )

for M0 x( )

for M x( )

Page 153: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 129

the driving variable vector (displacements or forces), obtained by solving the equilibrium

equations, does not necessarily lead to the best configuration of equilibrium, i.e., the one with

the least residual output vector (forces or displacements). Instead, a multiple of the

driving increment may give a better estimate of the equilibrium state, where the scaling factor,

or step length parameter is chosen so as to minimize some measure of residuals. Typically,

is optimum (Criesfield (1982)) when it makes the component of in the direction of

be zero.

In displacement based non-linear algorithms, stands for displacements , successively

corrected to satisfy equilibrium with the imposed force increment , and the residuals are

the out-of-balance forces . On the contrary, for force based schemes, such as the internal

element iterative algorithm considered in this study, consists of the forces , iteratively

adjusted in order to ensure compatibility with the desired displacements ; in this case

refer to the residual displacements violating kinematic compatibility.

Particularizing for the present force based scheme, for a given element internal iteration j, the

following expression holds

(3.47)

where the scalar is estimated by imposing

(3.48)

in which represents the work done by the increment of forces in the residual displace-

ments. Eq. (3.48) can also be interpreted as imposing the stationarity of the total potential

energy in the direction of .

Since and are known for iteration j, the scalar Eq. (3.48) requires only a one-dimen-

sional search of for which a trial-and-error procedure is often adopted. Each trial value of

(say ) involves the corresponding residual calculation , often called

“extra residual calculation” or simply “line search”. Different proposals of line search schemes

are available depending on more or less sophisticated algorithms for estimation of (Cries-

field (1982), Matheis and Strang (1979), Marques (1986)).

∆a

∆r β∆a

β( )

β ∆r ∆a

∆a ∆u

∆Q

∆Qr

∆a ∆Q

∆u ∆r

∆ur

Qj Qj-1 β∆Qj+=

β

ψ ∆Qj( )T

∆ur Qj-1 β∆Qj+( )⋅ 0= =

ψ ∆Qj

∆Qj

Qj-1 ∆Qj

β β

βh ∆urh Qj-1 βh∆Qj+( )

β

Page 154: Seismic Assessment of Reinforced Concrete Frame Structures

130 Chapter 3

In the present study the line search scheme adopted is proposed by Criesfield (1982), accord-

ing to which upper and lower values of are first sought, bounding the root of Eq. (3.48). The

solution is then estimated by successive linear interpolations between the lower and upper

bounds.

For , residuals are readily known as they correspond to the previ-

ous iteration (having given rise to , for ); additionally, the residual evaluation for

is always done as if no line search were used. Therefore, the two work residuals

(3.49)

are set first and the corresponding sign is checked. Taking as the basic value, the ratios

are calculated and plotted against , as shown in Figure 3.29.

Figure 3.29 Interpolation for line search scheme

If is negative, the lower and upper bounds are readily known as and

, respectively, and the new estimate is interpolated between them for ;

the residuals for are evaluated and a new value of (say ) is obtained. If is negative,

the upper bound is adjusted to and a new interpolation is performed (see Figure 3.29-

a)). Instead, if is positive, the lower bound becomes while the upper one is kept as

β

β 0= ∆ur Qj-1( ) ∆urj-1=

∆Qj j 2≥

β 1=

ψ0 ∆Qj( )T

∆ur Qj-1( )⋅ ∆Qj( )T

∆urj-1⋅= = for β 0=

ψ1 ∆Qj( )T

∆ur Qj-1 ∆Qj+( )⋅ ∆Qj( )T

∆urj⋅= = for β 1=

ψ0

ρ ψ ψ0⁄= β

a) Lower bound (βL ) unchanged

β1 1=β2

ρ1

ρ2

β3

ρ3 3

21

StopLine Search

Tolsr

-Tolsr

ρ0 1=

ρ ψψ0------=

1

b) Lower bound (βL ) shifted

β1 1=

β2ρ1

ρ2

β3

ρ33

2

1

StopLine Search

Tolsr

-Tolsr

ρ0 1=

ρ ψψ0------=

1

βL βLβU βUβU βL

ρ1 βL 0=

βU β1 1= = β2 ρ 0=

β2 ρ ρ2 ρ2

βU β2=

ρ2 βL β2=

Page 155: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 131

(see Figure 3.29-b)), allowing a new trial to be done by interpolation using the fol-

lowing expression

(3.50)

The process stops when is lower than a pre-defined tolerance TOLSR less than unity. This

tolerance depends on the characteristics and the powerfulness of the basic iterative procedure,

and states how far the line search is allowed to go. Usually some slackness is considered ade-

quate for line searches in conjunction with N-R or modified N-R algorithms (Criesfield

(1982)), which leads to TOLSR values set within 0.3-0.5 (Marques (1986)).

If is positive, an extrapolation process has to be set first to find a pair of and bound-

ing the zero of . In order to avoid dangerous extrapolations (which could be directly done

using Eq. (3.50)), the step is shifted forward, i.e., for and , residuals are

evaluated (only for is needed) and the ratio is checked for negative value. The process is

repeated until lower and upper bounds ( and ) are found (thus, for opposite signs of

and ), such that interpolation can be performed as stated before.

Once an acceptable value of (say ) is found, the corresponding residuals are taken as the

displacement residuals for iteration j, i.e., , and the state of all sec-

tions and the element stiffness matrix are updated accordingly. The next iteration j+1 is then

set up for application of corrective forces and new line searches may

eventually take place.

3.7 Summary of the non-linear algorithm

3.7.1 General structure

The element state determination as described in the previous section is nested within two outer

processes, namely one for the incremental sequence of load application and another for the

Newton-Raphson iterative scheme within each load step.

The incremental scheme for load application corresponds to the usually assumed loading his-

tory division into load increments or time steps, arising from either static or dynamic calcula-

βU β1=

β βLβU βL–( )

1 ρU ρL⁄–-------------------------+=

ρ

ρ1 βL βU

ρ

β1 1= β2 β1 1+=

β2 ρ2

βL βU ρL

ρU

β βa

∆urj ∆ur Qj-1 βa∆Qj+( )=

∆Qj+1 Kj-1 ∆urj–( )=

Page 156: Seismic Assessment of Reinforced Concrete Frame Structures

132 Chapter 3

tion. It follows the well known general non-linear scheme recalled in Appendix E, where the

Newton-Raphson (or external) iterative scheme refers to each step beginning as discussed in

2.4.4.4; thus, the corresponding expressions to be referred next are included in that Appendix.

Both the incremental sequence and the N-R scheme are included in the flow chart of Figure

3.30, the former corresponding to the outer loop over the load step k, and the latter being asso-

ciated with the loop for iteration n.

A unique scheme is adopted for both static and dynamic cases, but minor adaptations are high-

lighted where appropriate. It starts by computing the initial structure tangent stiffness ,

followed by the loop for step k, where the corresponding external load vector is first set

up and, if required, corrected for dynamic behaviour in task 2. This correction corresponds to

Eq. (E.12) in Appendix E, where (standing for ) is affected with contributions of iner-

tia and damping forces from the previous step (i.e., , according to

the Appendix E notation); additionally, the dynamic contribution to the effective stiffness

matrix is set up according to Eq. (E.18) as .

The structure stiffness matrix may be updated to the tangent one at each step beginning (see

task 3), depending on the adopted type of N-R scheme (standard or modified).

The initialization for the external iterative scheme is performed in task 4, where the increment

of nodal forces is set up according to Eq. (E.21) as the difference between the exter-

nal load vector and the nodal resultant force vector equivalent to the internal

stress state at the end of the previous step. In the same task, the stiffness matrix may be

selected as the tangent one, depending on the desired N-R scheme.

Task 5 accounts for the contributions of eventual dynamic behaviour to the force increment

and to the effective stiffness matrix; these two corrections, indeed correspond to the last term

of Eq. (E.19) and to the first of Eq. (E.18), respectively.

For a given iteration n, the displacement field increment is first updated (task 6), cor-

responding to Eqs. (E.14) and (E.20), and then transferred into an operator which performs the

state determination of all elements in the structure. The concept of operator is related with the

organization of CASTEM2000 (CEA (1990)) that will be addressed in 4.2; however, the fol-

lowing flowchart steps are described by reflecting already this basic philosophy of the code.

KG( )0

Qextk( )

qk Qextk

Qdynk M ak 1–⋅ C vk 1–⋅+=

KD 1 β∆t2( )⁄( )M γ β∆t( )⁄( )C+=

∆QG( )n=1

Qextk QG( )

k-1

K*G

∆uG( )n

Page 157: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 133

Figure 3.30 Flow chart for structure state determination

State determination of all elements

ielem = 1, number of elements

Set input: ∆uE( )n

∆Qf( )0

V( )0

State determination of element ielem

Output: Qf( )n

Vn

Set k 1= and assemble initial structure tangent stiffness matrix: KG( )0

Define k-th load vector: Qextk

If required: Update new structure tangent stiffness matrix: KG( )k

Set n 1=

initialize: ∆uG( )0

0= ∆QG( )n=1

Qextk QG( )

k-1–= K*

G KG( )k or KG( )

0=

Update: ∆uG( )n

∆uG( )n-1

K*G[ ]

1–∆QG( )

n⋅+=

Assemble resisting forces: QE( )n

Qf( )n

QG( )n

Element level Global assemblage

Compute structure unbalanced forces: ∆QuG( )

nQext

k QG( )n

–=

If required: Update structure tangent stiffness matrix

K*G KG( )

n=KE( )

nVn KG( )

n

Element level Global assemblageKn

sufficiently small ?∆QuG( )

n No∆QG( )

n+1∆Qu

G( )n

=

Next n

Yes

Next k

1

2

Task:

5

3

4

8

7

9

10

11

For dynamic behaviour, correct Qextk Qext

k Qdynk-1+= and define matrix KD

and

For dynamic behaviour correct ∆QG( )n

∆QG( )n

KD ∆uG( )n-1

⋅–=

K*G K*

G KD+=

;

6

Page 158: Seismic Assessment of Reinforced Concrete Frame Structures

134 Chapter 3

Besides the displacement field, the referred operator for task 7 requires also, as main input, the

field of element internal forces (assumed here as the forces at the flexible element end sec-

tions) and the field of element internal variables describing the element state, both fields

corresponding to the step beginning. The operator output consists of the updated fields of

forces and internal variables at the end of iteration n.

Since the element state determination contains the most innovative features in this work, this

summary will focus mainly on that, for a single N-R iteration n. However, for clarity sake, the

rest of the outer processes is still described before.

After completion of the element state determination, task 8 is accomplished by another opera-

tor which transforms the element internal forces into their equivalents in the total

element (with rigid lengths) reference system parallel to the global one, by means of transfor-

mations described in 2.4.2; these forces are then assembled to give the structure resisting force

vector .

Unbalanced forces are computed as described in task 9 of Figure 3.30 and, if required,

the tangent stiffness matrix may be updated in task 10 for the present iteration by means of

another operator. Based on the field of element internal variables , this operator performs

the calculation of each element stiffness matrix in its reduced space (without rigid body

modes) which then leads to in the total element global space by successive transforma-

tions. The global structure tangent stiffness matrix is then obtained by assembling the

element matrices . The adoption of for performing the next N-R iteration, again

depends on the choice made for the N-R scheme (standard or modified).

Convergence is checked in task 11 for sufficiently small forces and, in case it is not

reached, a new iteration n+1 is enforced for . When convergence is

reached, the next step k+1 is set up and a new increment application starts.

3.7.2 Element state determination

As shown in Figure 3.30 the input for each element state determination, for the n-th N-R itera-

tion, consists of the vectors , and referring, respectively, to the current itera-

tive increment of element displacements in its global reference system, to the flexible element

end section forces and to the element internal variables at the step beginning. Then, for each

Qf

V( )

Qf( )n

Vn

Qf( )n

QE( )n

QG( )n

∆QuG( )

n

Vn

Kn

KE( )n

KG( )n

KE( )n

KG( )n

∆QuG( )

n

∆QG( )n 1+

∆QuG( )

n=

∆uE( )n

Qf( )0

V0

Page 159: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 135

element, the n-th iteration develops as follows:

Step (1) Transform displacements due to reference system rotation .

Transform to the flexible element space (see Eq. (2.9)).

Extract rigid body modes (see Eq. (2.13)).

Split into linear and non-linear components , where

stands for a (6x1) dimension vector whose unique non-zero components are the

end section rotations in the non-linear bending plane.

Step (2) Extract reduced space initial forces , which can be deducted from

Eqs. (2.11) and (2.12) showing that components in the reduced space are not

affected by the transformation to the complete space.

Split into linear and non-linear force components , where

stands also for a (6x1) dimension vector containing only as non-zero components

the end section moments in the non-linear bending plane.

Initialize the increment force vector , such that at the

end of iteration n.

Step (3) Initialize the element reduced space flexibility and stiffness matrices, respec-

tively, and , based on the internal variables .

Split into linear and non-linear component contributions .

Step (4) If element loads p exist and the load increment is the first, compute the elastic

displacements due to p and split into contributions from linear and non-linear

components .

If loads p do not exist or the increment is not the first, set and .

Step (5) Start the element state determination for and initialize counter

for splitting according to events in the plastic zones.

Step (6) Initialize element internal state (for the step beginning):

- set up moving sections ( , , and );

- set up flexibility and curvature distributions associated with the

moment one .

Step (7) Set up element integration zones and compute by integration of , duly

corrected with the contribution of elastic shear distortions .

Step (8) If set .

∆uE( )n

∆ue( )n

∆ue( )n

∆uf( )n

∆uf( )n

∆un→

∆un ∆ulinn ∆θn+= ∆θn

Qf( )0

Q0→

Q0 Qlin0 M0+= M0

∆Qn 0= Qn Q0 ∆Qn+=

F0 K0 F0[ ]1–

= V0

F0 F0 Flin Fθ0+=

up

up uplinθp+=

up 0= θp 0=

∆θ ∆θn=

ib 1= ∆θ

Yi0 Ci

0 Ci+20 Oi

0

f0 x( ) ϕ0 x( )

M0 x( )

θ0 ϕ0 x( )

β0 x( )

ib 1> θp 0=

Page 160: Seismic Assessment of Reinforced Concrete Frame Structures

136 Chapter 3

Search for events in plastic zone boundary sections and associated with

the application of end section moments and the distribu-

tion of element loads. Set .

If required, split according to the maximum number of events found in any

of those sections, such that at most one event will be found per section; thus,

Step (9) Initialize internal iteration counter and increment .

Step (10) Initialize Line Search counter and save .

Step (11) If set .

Compute iterative increment of end section moments with the previous stiffness

matrix .

Step (12) If , initialize or update variables and residuals of the Line Search process.

Step (13) Search for events in and associated with the application of and sub-

divide according to the event which occurs first, i.e. for the least reduction

factor found among sections having events . Thus,

Step (14) Update end section moments and the distribution along the element,

where the contribution of the element applied loads p (in case it exists) is always

included by means of the adequate expressions of (or ) as included

in Appendix C:

Step (15) Search for new moving sections , , and for the present diagram

, such that

Ei Yi0

∆M K0 ∆θ θp–( )⋅=

ib ib 1+=

∆θ

∆θ∆θb (basic)

∆θr ∆θ ∆θb–( ) (remaining) =⎝⎜⎛

j 1= ∆θj=1 ∆θb=

iLS 1= ∆θLS ∆θj=

j 1> θp 0=

∆Mj Kj-1 ∆θj θp–( )⋅=

j 1>

Ei Yi0 ∆Mj

∆θj

r( ) 0 r 1≤<( )

∆θj r∆θj= set ∆θj ∆θj=→

∆Mj Kj-1 r∆θj θp–( )⋅= set ∆Mj ∆Mj=→

Mj Mj x( )

Sp x( ) Spsx( )

Mj Mj-1 ∆Mj+=

Mj x( ) Mj-1 x( ) ∆Mj x( )+=

Yij Ci

j Ci+2j Oi

j

Mj x( )

sYi

j sYi

0≥ sCi

j sCi

0≥ and sCi+2

j sCi+2

0≤

Page 161: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 137

Define the corresponding derivatives for the additional flexibility terms, depend-

ing upon eventual further progression of cracking sections and taking into

account possible full-cracking along one (both) element part(s).

Step (16) Perform the state determination of fixed sections ( and H) and of . Due to

the increment reduction imposed by the event-to-event procedure and to the

multi-linear character of the M-ϕ curve, the total curvature in sections

and is given by .

In the section (or sections) having generated the conditioning event, the flexibil-

ity is updated to after the event, whereas for the remaining sections the

flexibility is kept .

For the section H, the M-ϕ diagram is one-to-one and the total curvature and

the flexibility are obtained directly from .

Step (17) Perform the state determination of the moving sections , , and , for

which and are directly obtained from the with their one-to-one M-ϕ

diagrams.

Generate the additional flexibility terms due to eventual progression of and/or

for the current iteration.

Step (18) Update integration zones, based on the activated moving sections.

Step (19) Integrate the flexibility diagram to obtain the flexibility terms of associ-

ated with the non-linear component of deformation, and include the additional

contributions due to moving sections.

Accumulate with linear contributions and invert to obtain the stiffness matrix

Step (20) Integrate the total curvature diagram to obtain total rotations and

include the contributions due to elastic shear distortion.

Compute displacement residuals (necessary to assure equilibrium) given by:

Step (21) Check displacement convergence:

If is verified according to 3.6.8, proceed to Step (23).

Ei Yi0

ϕ…j Ei

Yi0 ϕ…

j ϕ…j-1 f…

j-1∆M…j+=

f…j-1 f…

j

f…j f…

j-1=

ϕHj

fHj MH

j

Yij Ci

j Ci+2j Oi

j

ϕ…j f…

j M…j

Cij

Ci+2j

fj x( ) Fθj

Fj Flin Fθj+= Kj Fj[ ]

1–=

ϕj x( ) θj

δθj

δθj θj θ0 ∆θb––=

max δθ1 δθ2( , ) θtoler<

Page 162: Seismic Assessment of Reinforced Concrete Frame Structures

138 Chapter 3

In case convergence is not reached:

- if , set and restart in Step (10) for next iteration ;

- if proceed to Step (22).

Step (22) Control line search residuals:

If line search tolerances are verified according to 3.6.9, accept residuals and

proceed for the next iteration restarting in Step (10) with .

If tolerances are not satisfied, search for new scaling factor , set new iteration

scaled increment and restart new line search trial in

Step (11).

Step (23) Convergence is reached for a state of internal equilibrium in the element, corre-

sponding to the sub-increment (possibly the total increment). Update ele-

ment state for the end of :

Step (24) Check if there is still a remaining sub-increment to be applied:

If , set and restart in Step (8).

If , no remaining sub-increment exists and proceed in Step (25).

Step (25) Complete with the contribution of elastic forces, which can be simply cal-

culated by due to the uncoupling of force (or displacement) compo-

nents contributing to different deformation directions. Therefore:

Step (26) Transform force increment to the complete space of the flexible element, eventu-

ally including contribution of element forces p: (see Eq. (2.11));

Compute the final forces in the flexible element reference system for the n-th

Newton-Raphson iteration, given by , and the final out-

j 1=( ) ∆θj+1 δθj–= j 1+( )

j 1>( )

δθj

j 1+( ) ∆θj+1 δθj–=

β

∆θj β∆θLS= iLS iLS 1+=

∆θb

∆θb

θnew0 θ0 ∆θb+= → set θ0 θnew

0=

∆Qnewn ∆Qn Mj=converg M0–( )+= → set

M0 Mj=converg=

∆Qn ∆Qnewn=⎝

⎜⎛

setV0 Vj=converg=

K0 Kj=converg=⎝⎜⎛

∆θr

∆θr 0≠ ∆θ ∆θr=

∆θr 0=

∆Qn

K0 ∆ulinn⋅( )

∆Qtotn ∆Qn K0 ∆ulin

n⋅( )+= → set∆Qn ∆Qtot

n=

Vn Vj=converg=⎝⎜⎛

∆Qn ∆Qf( )n

Qf( )n

Qf( )0

∆Qf( )n

+=

Page 163: Seismic Assessment of Reinforced Concrete Frame Structures

FLEXIBILITY BASED ELEMENT WITH MULTI-LINEAR GLOBAL SECTION MODEL 139

put of the element state determination is available ( and ) corresponding

to the element displacements .

The above described sequence of steps has been included in the CASTEM2000 operator which

performs the state determination of all structure elements, and, indeed, constitutes the most

significant contribution to that object-oriented and general purpose computer code. However,

several other adaptations had to be done in other operators (e.g., the one for tangent matrix cal-

culation) in order to cope with the present element specific features. Comments on such adap-

tations and further details about the code structure are included in Chapter 4.

3.8 Concluding remarks

In the present chapter the general flexibility formulation has been particularized for the case

where the global section behaviour is described by a multi-linear model based on a trilinear

skeleton curve.

Use is made of the stiffness variations at cracking and yielding points of the trilinear curve, in

order to define the so-called cracking and yielding sections. These sections move along the ele-

ment and allow its division into several zones with distinct behaviour. Thus, control sections

are considered of fixed type (the member-end and the mid-span ones) and of moving type (the

cracking, yielding and null-moment sections), which permit a continuous updating of the ele-

ment flexibility distribution during the response. Basically, the element becomes subdivided

into a number of zones, viz the plastic (adjacent to the yielded end sections), the cracked and

the uncracked zones, the two latter developing in between the plastic ones according to the

bending moment distribution.

Modifications have been introduced in an existing behaviour law, mainly in the post-cracking

range in order to make possible the control of cracked zones by means of an efficient, yet

approximate way. Specifically, a transition from uncracked to cracked behaviour is considered

localized in the section where cracking is incipient, in such a way that sections in the cracked

zones keep a linear elastic behaviour, though with reduced stiffness. The major advantage is

that the fully-cracked stiffness is progressively introduced in the element according to the

actual loading; however, at the present development stage, it does not allow energy dissipation

by hysteresis in cracked zones.

Qf( )n

Vn

uE( )n

uE( )0

∆uE( )n

+=

Page 164: Seismic Assessment of Reinforced Concrete Frame Structures

140 Chapter 3

The element state determination is performed by means of an internal iterative scheme where

corrections to element end-forces are successively made by elimination of element displace-

ment residuals and the element tangent flexibility (and stiffness) matrix is continuously

updated; the calculation of this matrix and those residuals constitute the major and most inno-

vative tasks in the developed scheme.

Due to the specific character of moving sections, their state determination is readily performed

once their nature and location is defined for a given distribution of moments.

The element flexibility is determined according to the instantaneous location of control sec-

tions where the flexibility is known from the corresponding model; between control sections

the flexibility is assumed linearly distributed, which is coherent with the uncracked and

cracked section models and, under certain circumstances, also with the model of yielded sec-

tions. The moving character of cracking sections, where a curvature discontinuity is assumed

to account for the uncracked/cracked behaviour transition, has been found to introduce special

contributions to the element flexibility matrix which have to be duly included in order to suc-

cessfully account for that transition.

The element residual displacements are calculated by an alternative scheme rather than that of

the general flexibility formulation; it is based on integration of total deformations instead of

residual ones, since it has been found more compatible with the control of moving sections.

The control of plastic end zones has been carefully detailed. Each of those zones is subdivided

into one fixed-length part and another with variable length (during the application of a given

load increment), in order to minimize possible inaccuracies resulting from the fact that the

plastic zone behaviour is just based on the end section and yielding section behaviour. In the

same line, an event-to-event scheme is used inside the element iterative process.

Convergence difficulties related with the uncracked/cracked behaviour transition required the

adoption of a line-search scheme, which, being internal to the element iterations, has proved to

be quite efficient.

Finally, the steps of the element iterative process are summarized for a given iteration of the

global non-linear algorithm of the structure, and the interface between both schemes is high-

lighted in a flowchart illustrating the main tasks of the structure state determination process.

Page 165: Seismic Assessment of Reinforced Concrete Frame Structures

Chapter 4

NUMERICAL IMPLEMENTATION,

AUXILIARY TOOLS AND VALIDATION

4.1 Introduction.

The flexibility formulation as described in the previous chapter has been implemented in the

general purpose computer code CASTEM2000 (CEA (1990)), and the main implementation

related topics are addressed in 4.2.

In view of the particular features of the code, indeed rather different from traditional specific

purpose computer codes, a brief description of CASTEM2000 basics is first provided in 4.2.1

and complemented with a rather simple example of structural analysis, in order to present the

user interface environment. The basic code structure and the intervention levels for new code

improvements are addressed. Subsequently, the most significant implementation needs are

identified within the framework of previously introduced concepts such as objects, commands,

operators and procedures. In particular, the major implementation topics related with the flexi-

bility element formulation are referred in 4.2.2 and the operators requiring the most relevant

modifications are briefly presented in order to point-out the performed interventions.

A new algorithm is presented in 4.2.3 for the definition of trilinear approximations of moment-

curvature relationships for rectangular and T-shape reinforced concrete sections. This pre-

processing task is described with some detail and particular emphasis is put on a unified proc-

ess for definition of turning points in the non-linear behaviour range; the implementation of

this auxiliary tool in CASTEM2000 has been cast in the form of a new operator.

Some validation tests at the single element level are included in 4.3 for experimentally tested

Page 166: Seismic Assessment of Reinforced Concrete Frame Structures

142 Chapter 4

members under monotonic and cyclic loading conditions. Rather than aiming at an exhaustive

process of model validation against experimental results, the tests presented allow to find-out

the expectable quality of numerical results and to identify model limitations that must be kept

in mind when analysing complete structures.

4.2 Implementation in the computer code CASTEM2000

4.2.1 Basics of CASTEM2000 and main implementation needs

CASTEM2000 is a multi-purpose finite element based computer code for structural analysis,

initially developed by the “Commissariat à l´Energie Atomique” (CEA) in the framework of

structural mechanics research (CEA (1990)). The need of treating several types of problems

based on different formulations (solid and fluid mechanics as well as thermal processes),

required the development of a high level tool of analysis based on a unified and powerful tech-

nique such as the finite element method.

Aiming at a unified way of handling different problems, the code has been structured follow-

ing the, nowadays increasingly adopted, object-oriented technique of programming. It is based

on a specifically developed high level language GIBIANE (or simply GIBI) (CEA (1990))

consisting on a wide set of commands and operators used to control and define the program

flow by object manipulation in a specific environment or shell.

The object-oriented features of CASTEM2000 lead to a high level of versatility and flexibility

in the sense that it can be adjusted to the particular problem to be solved. In contrast with clas-

sical codes designed for the analysis of certain well-defined type of problems, to which spe-

cific cases have to be adjusted, CASTEM2000 allows the user to build-up the program flow by

himself, to follow the analysis task-by-task, to modify the task sequence, to re-define tasks and

to check their outputs, in a word, to adapt it to his own needs.

The macrolanguage GIBI permits to define the usual operations characteristic of finite element

analysis, by means of simple instructions involving commands or operators acting on input

objects and, possibly, generating new output objects. A typical case of such operations is the

stiffness matrix assemblage: previously defined element mesh and fields of material properties

and element characteristics are provided as input objects for the operator which performs ele-

ment stiffness computation and assemblage, giving the global stiffness matrix as a new object.

Page 167: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 143

Objects are defined as pieces of information grouped according to specific and well defined

rules characterizing the object type. Some of the main object types are briefly referred next;

inside brackets it is included the original french name used in the code for the object type:

• geometric types, specifically node (POINT) and mesh (MAILLAGE);

• scalar or vector field types, defined on the nodes or the elements, namely the nodal field

type (CHPOINT) and the element field type (MCHAML), respectively; the former stands

for example for displacements, forces or temperatures, whereas the latter may refer to mate-

rial properties, geometric characteristics or internal stress/strain components, defined on all

the element integration points;

• model type (MODELE), which includes references to the formulation associated with the

finite elements in a given mesh and to the constitutive behaviour model;

• stiffness type (RIGIDITE), consisting of the material stiffness or mass matrices associated

with a given mesh, or the stiffness corresponding to imposed boundary conditions;

• loading type (CHARGEMENT), containing information about a given force or displace-

ment field representing the load, and, possibly, the time description of the load process;

• integer (ENTIER), real (FLOTTANT) and string (MOT) types, standing for single numeri-

cal/string constants or variables;

• table type (TABLE), containing a set of objects of any type, identified by numerical indices

or strings.

Commands (DIRECTIVES) and operators (OPERATEURS) are used to perform operations on

input objects allowing to manipulate them (by modifying them or not) in the first case, and to

generate new objects in the second case. The available commands and operators in the GIBI

language can cover a wide range of purposes as different as:

• general operations, namely direct generation of objects (e.g. copying), object management

(e.g. listing, deleting), mathematical operations (e.g. arithmetical, logical, trigonometrical),

modification and extraction of data from objects (e.g. maximum value of a scalar field) and

flow control operations (e.g. loops, “if-then-else” type blocks, etc.);

• preparation of the analysis model, viz, geometric mesh generation, model definition (formu-

lation, type of element, type of behaviour model), material properties, element characteris-

tics, boundary conditions (fixing supports, imposed displacements, relations between

degrees of freedom) and loading;

• solution of a discretized problem, namely definition and assemblage of stiffness and mass

Page 168: Seismic Assessment of Reinforced Concrete Frame Structures

144 Chapter 4

matrices, solution of linear equation systems and of eigensystems;

• analysis of results, consisting of post-processing computations (stress and strain fields,

reactions, etc) and of graphical visualization.

Commands and operators can be organized following a user-defined sequence of tasks in order

to perform the desired analysis. Such sequence constitutes the so-called GIBI input for

CASTEM2000 running sessions, either in interactive or batch mode.

An illustrative example of a GIBI input is given below for the very simply structure shown in

Figure 4.1, where uniform elastic material properties and cross-section characteristics are

assumed as indicated.

Figure 4.1 Illustrative example for GIBIANE input

In the following input list, commands and operators are highlighted with boldface italic letters,

upper case words stand for the names of expected arguments and lower case, normal typeface

letters refer to “values” of such arguments.

1/ opti DIME 3 ELEM seg2;

2/ p0 = 0.0 0.0 0.0;p1 = 0.0 0.0 4.0;p2 = 3.0 0.0 4.0;

3/ c1 = droi 1 p0 p1;b1 = droi 1 p1 p2;

4/ ll_1 = c1 et b1;

5/ mo_1 = modl ll_1 mecanique elastique poutre;

6/ ma_1 = matr mo_1 YOUN 20.e6 NU 0.3;

7/ ca_1 = carb mo_1 SECT 0.36 INRY 0.0108 INRZ 0.0108 TORS 0.0216;

8/ mc_1 = ma_1 et ca_1;

9/ bll1 = bloq rota depl p0;

10/ bll2 = (bloq UY ll_1) et (bloq RX ll_1) et (bloq RZ ll_1);

11/ ri_ll = rigi mo_1 mc_1;

4.0

m

F1 = 100 kN

3.0 m

y

x

z

c1

b1

E = 20x106 kPaMaterial properties

Section Characteristics

ν = 0.3

A = 0.36 m2

Iy = Iz = 0.0108 m4

Ix = 0.0216 m4p0

p1 p2

Page 169: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 145

12/ ritot = ri_ll et bll1 et bll2;

13/ f1 = forc FZ -100.0 p2;

14/ d1 = reso ri_ll f1;

15/ sd1 = defo ll_1 d1;

16/ pv = 0.0 -1.e8 0.0;

17/ trac pv sd1;

18/ s1 = sigm mc_1 d1;

19/ list s1;

The input line 1/ stands for the command opti where the use of the tridimensional space is

declared by the argument DIME with the value 3; the type of geometric supports for the finite

elements is set by the argument ELEM for the element named seg2 consisting of a two-node lin-

ear segment. The coordinates of nodes p0, p1 and p2 are set in lines 2/ and the geometric ele-

ments c1 and b1 (for the column and the beam, respectively) are defined in lines 3/ by the

operator droi stating that 1 single element is considered between the extreme nodes. The total

mesh ll_1 is then obtained in line 4/ by the operator et which performs the concatenation of the

partial meshes referring to the column c1 and the beam b1.

The model type object mo_1 is obtained in line 5/ from the operator modl which associates the

mesh ll_1 with the mechanic formulation (mecanique), the elastic behaviour (elastique) and

beam finite elements named poutre, consisting in the classical Bernoulli type element. Material

properties (assumed uniform) are stored in the element field ma_1 generated by the operator

matr where the Young modulus YOUN and the Poisson ratio NU are declared in line 6/. Addi-

tionally, the operator carb allows to define in line 7/ the element characteristics (not obtainable

from the mesh), namely the cross-section area and moments of inertia, stored in the element

field ca_1 as components labelled by their argument names (SECT, INRY, INRZ, and TORS). The

total set of material properties and characteristics is grouped in the field mc_1 by the operator

et in line 8/.

Boundary conditions associated with the fixed support are set in line 9/ by the operator bloq,

which, under the keywords rota and depl completely blocks rotations and displacements of the

node p0; the result consists of a stiffness type object bll1. Similarly, the total mesh ll_1 is pre-

vented to have displacements UY (in the yy direction) and rotations RX and RZ, as enforced by

the operator bloq in line 10/, where three boundary conditions are imposed and then concate-

nated in only one stiffness type object bll2.

Page 170: Seismic Assessment of Reinforced Concrete Frame Structures

146 Chapter 4

The stiffness type object ri_ll due to the elastic material behaviour is constructed in line 11/ by

the operator rigi whose arguments are the model mo_1 and the element field mc_1 of properties

and characteristics. The total structure stiffness ritot is then obtained by concatenation of the

relevant contributions in line 12/, namely the material stiffness ri_ll and the boundary condition

ones bll1 and bll2.

The applied force field f1 is defined in line 13/ by the operator forc imposing the value -100.0

for the force component FZ in the point p2.

The linear system with stiffness ritot and subjected to the force field f1 is solved by the operator

reso (a linear equation system solver) in line 14/, leading to the displacement field d1.

The phase of result analysis starts in line 15/, with the generation of the deformed shape sd1 (a

mesh type object) by the operator defo upon the original mesh ll_1 and the displacement field

d1. This deformed shape sd1 can be visualized in line 17/ by means of the command trac

requiring a viewpoint pv (previously defined in line 16/).

The operator sigm is used in line 18/ to generate the element field s1 containing the internal

stresses (in this case the end section forces), from the properties and characteristics mc_1 and

the displacement field d1; the contents of s1 is then listed in line 19/ as the result of the com-

mand list.

The above given example aims just at a general but illustrative perspective of how

CASTEM2000 works and, for complex problems, it may appear somewhat cumbersome. In

such cases the use of procedures becomes extremely advantageous; procedures are sequences

of operators cast in independent GIBI segments and acting as higher level operators to accom-

plish well defined purposes. Actually, procedures play the same role as subroutines do in com-

mon programming languages, allowing to perform rather standard and repetitive sets of tasks.

Typical examples of procedures are the implementation of the Newmark method for the inte-

gration of dynamic equilibrium equations or the non-linear incremental analysis using the

Newton-Raphson algorithm.

CASTEM2000 provides a set of built-in procedures to accomplish some usual tasks in struc-

tural analysis which cannot be handled by a single operator, but other procedures can be easily

designed and implemented by the user.

Page 171: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 147

Thus, concerning tool implementation or improvement, the code offers two possible ways, viz:

• Development of procedures, written in GIBI (thus, strictly relying upon existing operators)

when the envisaged tasks do not involve new elements, models or formulations; in these

conditions, this is a low cost option from the implementation standpoint since it allows a

very fast development and on-line testing, without the need of modifications at the basic

CASTEM2000 software level.

• Development of new operators, based on existing and new subroutines constituting the code

source software; new operators are required when not yet available elements, models or for-

mulations are to be incorporated, and may be appropriate for efficiency purposes when cer-

tain algorithms, despite also implementable at GIBI level, would lead to cumbersome and

computationally heavy procedures; the implementation cost of operators is obviously

higher than that of procedures, since it requires a more in-depth knowledge of the code data

structure in order to provide the adequate operator interface.

In the present work, developments have been performed at both procedure and operator levels,

although the main contribution consisted of new improvements on existing operators in order

to incorporate the proposed flexibility based element; the most significant interventions

required for these developments refer essentially to:

• the inclusion of the new flexibility based element, for which all the affected existing opera-

tors needed to be adapted;

• the development of a new operator for the definition of skeleton curves for RC global sec-

tion modelling, indeed a quite cumbersome but crucial pre-processing task;

• slight adaptations of the general procedure for the non-linear static and dynamic analysis,

which essentially follows the scheme described in Appendix E and the flowchart included

in Figure 3.30;

• several post-processing procedures for result analysis and visualization of frame structures.

The two first interventions, being related with operator development, require a few comments

about some key issues on CASTEM2000 organization concerning finite element formulations

and models, as well as on the most relevant features of internal programming.

As far as formulations and models are concerned, the following aspects are highlighted:

• Finite elements are treated by distinguishing the geometric support and the underlying for-

mulation; therefore,

Page 172: Seismic Assessment of Reinforced Concrete Frame Structures

148 Chapter 4

- various types of geometric supports can be adopted simultaneously in the same structure;

- each geometric support can be assigned more than one finite element formulation,

depending on the problem type (mechanic, thermic, ...) and on specific characteristics

related with the internal distribution of the unknown variables; a typical example is the

two-node linear segment used as geometric support of both the truss element and the

Bernoulli or the Timoshenko beam elements.

• Each finite element can be used with different behaviour models, provided the necessary

consistency is assured between the element formulation and the model; an example of that

is the two-node Timoshenko finite element, which can be used either with global section

behaviour laws (for both moment-curvature and shear-distortion components) or with a

fibre discretization approach where the behaviour is controlled at each fibre level (Guedes

et al. (1994)).

• In a given structure, several zones can be defined where different combinations of one sin-

gle finite element type with one element model are considered; thus, there can be as many

zones as the number of different combinations of elements and models.

It is apparent that such features provide a high degree of versatility to handle very different

types of problems, which is further assisted by a powerful set of tools for mesh generation.

Concerning the programming features, it is worth mentioning that the source code is written in

an extended FORTRAN77 language, the so called ESOPE language, which includes a few

additional instructions for management of data structures. Basically, arrays of data are grouped

into larger data segments which are initialized, activated, de-activated or suppressed according

to the code flow needs. The required data for subroutines to perform their tasks, is made avail-

able by some of those extra instructions; once the data is no more needed, other instructions are

used to make it unavailable again.

Each operator is supported by a driver, i.e., a subroutine (written in ESOPE) where the input

and output objects (fields, models, tables, etc) are decoded into segment-based data structures

managed by ESOPE instructions. The data is then transferred to lower level subroutines where

the structural calculations are performed; typically, the lowest level subroutines just handle

data in the traditional way of FORTRAN, which renders more transparent and easy the core of

implementations where the basic structure-related operations are performed.

Page 173: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 149

This means that, when new operators do not need to be developed, the implementation is quite

straightforwardly accomplished by normal FORTRAN subroutines, requiring only a careful

identification of the data to be transferred from/to the upper level CASTEM2000 subroutines

more directly related with the operator driver.

In order to cope with the above referred structure subdivision into zones, subroutines of opera-

tors dealing with objects extended over the whole structure (or part of it), typically perform the

three following nested loops:

• the outer loop over the total number of zones in the structure,

• another loop extended over all the elements in each zone and

• finally, the inner loop over all integration (Gauss) points in each element.

In some operators and for certain type of elements, eventually the inner loop may be skipped,

as in the case of the elastic stiffness evaluation of beam or truss elements where the result can

be directly derived without considering the integration points.

4.2.2 Flexibility based element implementations

The element developed in the present work required the intervention in several operators,

where a new finite element formulation and a modified (new) model had to be included.

Among the already available elements in CASTEM2000, the Bernoulli beam element is the

closest to the flexibility based one. Therefore, despite their different characteristics, the imple-

mentation scheme for the flexibility element followed that of Bernoulli element as close as

possible. With this in mind, the main similarities and differences, with respect to the Bernoulli

element implementation, are highlighted in the following comments:

• the same geometric support has been used, viz the two-node segment named seg2;

• the same internal generalized forces and deformations, as well as nodal forces and displace-

ments were considered (as defined in 2.4.1);

• a new formulation has been defined and included because

- there are no deformation shape functions and the formulation must be cast consistently;

- an internal span section H is considered, bounding the two element parts likely to have

different properties, although uniform in each of them;

- rigid lengths are admitted in the element end zones;

Page 174: Seismic Assessment of Reinforced Concrete Frame Structures

150 Chapter 4

- the sections where generalized internal forces and deformations are controlled (thus,

playing the role of integration Gauss points in the classical finite elements) are located at

the end points of the flexible element, not necessarily coincident with the element nodes;

• a different non-linear model is considered, although based on the existing Takeda type

model for use with the Bernoulli element;

• no inner loop is performed over the integration points because the state determination is car-

ried out for each element as a unique entity following the steps described in 3.6 and 3.7.

Consequently, a new finite element named FLD1 and a new model designated by

TAKEMF_MOMY were added to the lists of available finite elements and models, respectively.

Additionally, a new type of formulation named FLEXIBIL was included in the formulation list.

Contrarily to the available beam elements, different material properties and element character-

istics can be assigned to its left and right parts. Thus, the element fields containing these data

may have different values in the two integration sections, which for the flexibility element are

the end sections of the flexible part but do not assume the usual role of integration section.

In terms of element characteristics, the same list as for the Bernoulli element was kept (namely,

the cross-sectional areas and the moments of inertia defined in Appendix A and a vector defin-

ing the orientation of the section axis system), to which the rigid lengths in each end zone and

the relative abscissas of the span section H have been added under the keynames LORG and

XH_L.

Concerning internal variables, while the Bernoulli element requires only variables for the con-

trol of sections fully governed by the behaviour model, in the flexibility element more infor-

mation has to be stored related with the yielding and cracking development in each element

part. Therefore, three new variables were added and given the keynames XYLD, XCRK and

XCRI, respectively standing for the yielding and cracking section abscissas , and ,

as defined in 3.4 and associated with each end section. Additionally, three other variables

named FLE1, FLE2 and COF2 were also included, related with the evolution of the element flex-

ibility and the curvature distributions. FLE1 and FLE2 stand, respectively, for the diagonal and

off-diagonal non-linear terms of the flexibility matrix as defined in 3.6.3 by Eq. (3.18)

(namely, FLE1 for and , and FLE2 for and ); COF2 is the quadratic term of the

curvature distribution in each plastic end zone as defined in 3.6.6.

sYisCi

sCi 2+

F33M F55

M F35M F53

M

Page 175: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 151

Operators where the most significant modifications had to be made are referred next, along

with a brief description of their syntax and purpose; the notation used for operators and objects

is the same of the example given in the previous section.

1/ Operator: rigi

Syntax: zrige = rigi zmodl zmatr

Purpose: calculation of the structure elastic stiffness with the model zmodl and the field of material

properties and element characteristics zmatr; the output is the stiffness type object zrige;

2/ Operator: sigm

Syntax: zsige = sigm zmodl zdepl zmatr

Purpose: calculation of the elastic element stresses (or internal generalized forces) associated with a

given displacement field zdepl; the output is the element field type object zsige;

3/ Operator: bsig

syntax: zforc = bsig zmodl zsigm zmatr

Purpose: to obtain and assemble the element nodal forces equivalent to a given internal stress distribu-

tion (zsigm); the output is the nodal field type object zforc;

4/ Operator: epsi

syntax: zepsi = epsi zmodl zdepl zmatr

Purpose: calculation of the element internal strains (or generalized displacements) for a given dis-

placement field zdepl; the output is the element field type object zepsi;

5/ Operator: ecou

syntax: zsig1 zvar1 zdpl1= ecou zmodl zsig0 zvar0 zdeps zddep zmatr ztabl

Purpose: to perform the state determination on non-linear systems (updating stresses and internal var-

iables) for a given increment of strains zdeps or displacements zddep imposed over an

equilibrium state characterized by stresses zsig0 and internal variables zvar0; further input

is required in the table type object ztabl containing additional information for the control of

the non-linear process; the output consists of the element field type objects zsig1, zvar1 and

zdpl1, respectively, the updated stresses, internal variables and plastic deformations (not rel-

evant for the present implementation);

6/ Operator: ktan

syntax: zrig1 = ktan zmodl zsig1 zvar1 zmatr

Purpose: calculation of the structure tangent stiffness for a state characterized by element stresses

zsig1 and internal variables zvar1; the output is the stiffness type object zrig1.

In the above referred operators, new subroutines had to be included inside the loop over the

elements in order to accomplish each operator task for the flexibility based element. Thus,

beyond the preparation of data to be transferred into/from the new subroutines, the following

Page 176: Seismic Assessment of Reinforced Concrete Frame Structures

152 Chapter 4

adaptations had to be made:

rigi a master subroutine (fldsti) was included (calling slave routines) to obtain the ele-

ment elastic stiffness matrix via the flexibility one according to Appendix

A, followed by inversion and by the transformations given by Eqs. (2.26) and (2.27),

and by appropriate reference system rotations;

sigi inclusion of a master subroutine (fldefe), and slave routines, where the elastic forces

are computed for given displacements , with the matrix obtained

as in the rigi operator; the transformation to forces is also performed according

to Eq. (2.11);

bsig inclusion of subroutine (fldbsg) to perform the transformation of forces to in

the element nodes (as expressed by Eq. (2.9)), followed by reference system rotation

to obtain in the global reference system;

epsi inclusion of subroutine (fldeps) to transform the element displacements , first by

reference system rotation to the element local axes, giving , and then to as

expressed by Eq. (2.9);

ecou a master subroutine (fldesd) was included to perform the element state determina-

tion for non-linear behaviour according to Chapter 3 and following the steps

described in 3.7; indeed it is the major intervention within this implementation,

where several slave subroutines are controlled, of which the main ones are:

fldisd, for preparing the input at the reduced space level (see steps 1 to 3 in the ele-

ment state determination sequence in 3.7) and

fldipe, for performing the iterative process in the element (corresponding to steps 4

to 26 of the same sequence);

ktan a master subroutine (fldstt) was included (calling slave routines) to obtain the ele-

ment tangent stiffness matrix via the flexibility one , where the non-linear

contributions are extracted from the internal variables; the linear contributions and

appropriate matrix transformations are handled as in the rigi operator.

The aforementioned interventions just provide an overview of the basic needs for the flexibil-

ity based element implementations. Naturally, the whole set of modifications and new code

segments actually developed went far beyond the simplified description made herein. How-

ever, further details are deemed unnecessary in the present work, since they can be found else-

where (Arêde and Pinto (1997)).

K( ) F( )

Q K u⋅=( ) u K

Qf

Qf Qe

QE

uE

ue uf

K( ) F( )

Page 177: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 153

4.2.3 Definition of skeleton curves for RC global section modelling

The present section deals with the definition of the so-called trilinear approximation of the

moment-curvature relationship for reinforced concrete sections under uniaxial bending condi-

tions. Figure 4.2 illustrates the typical shape of this relationship (dashed line curve) and the

corresponding trilinear approximation (solid line curve) defined by the usual turning points C

(cracking), Y (yielding) and U (ultimate), as referred in 3.3.

Figure 4.2 Typical moment-curvature diagram and trilinear approximation

It has already been mentioned that the M-ϕ curve, or just the turning points for a trilinear

approximation, can be obtained by a fibre-type analysis, since this constitutes a general tech-

nique for bending analysis of reinforced concrete sections. However, the fibre or layer discreti-

zation required at the section level may become computationaly rather heavy when large

number of sections have to be analysed.

Taking into account that most of the sections in frame structures reduce to rectangular and T-

shape ones, and making a systematization of the internal equilibrium analysis in such type of

sections, it is possible to build-up an algorithm through which the section state is assessed for

each turning point without the need of section discretization. Although based on an iterative

scheme (due to the piecewise non-linear behaviour models), the algorithm was found quite

efficient and very easy to use and, furthermore, it can be seen as an alternative way to trace-out

the M-ϕ curve in the post-yielding range.

4.2.3.1 Type of sections, notations and conventions

The following study focus on the most common cross-sections of structural elements in rein-

forced concrete buildings, viz rectangular for columns and T-sections for beams; the assumed

MMu

My

Mc

ϕϕuϕyϕc

C

YU

Page 178: Seismic Assessment of Reinforced Concrete Frame Structures

154 Chapter 4

general layout is shown in Figure 4.3 and the meaning of section characteristics is as follows:

• stand for the width and height of the section web;

• refer to the width and height of the section slab contribution;

• ; stand for the areas and depths of the bottom and top steel layers (main

reinforcement);

• ; ;... refer to the areas and depths of the interior steel layers (secondary

reinforcement);

• is the slab reinforcement, assumed included in ;

• is the concrete cover, measured from the concrete surface to the stirrups and assumed uni-

form along the stirrup perimeter.

Figure 4.3 Types of sections

The T-shape section can be assumed either in the position shown (flange upwards) or in the

reverse (flange downwards); in any case the meaning of characteristics is always as described

above.

Concerning the forces acting in the section and the internal generalized stresses and strains, the

following convention is adopted, along with the reference axis systems shown in Figure 4.4:

• acting forces ( and ) are applied at the geometric centre of the rectangular concrete

gross section with positive directions according to the reference axis system

passing through that point.

• internal generalized strains ( and ) are referred to the central principal axis system

, containing the section centroid for linear elastic behaviour or the neutral axis

for non-linear behaviour.

b h,

bs hs,

As1d1,( ) As2

d2,( )

As1

I d1I,( ) As2

I d2I,( ) nI

AsLAs2

c

ASLAS2

AS1

hs

bs

d1 h

d2

d1 h

b

d2I

d1I

AS2

I

AS1

I

a) Rectangular section b) T-shape section

c

a2

a1

AS1

d2

AS2

b

Nx My

Cg( )

xg yg zg, ,( )

εx ϕy

x y z, ,( ) G

Page 179: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 155

Figure 4.4 Reference axis system convention

Since only the bending component is of interest, in the following the subscripts refer-

ring to axes are omitted and, thus, , , and are simply denoted by N, M, and .

4.2.3.2 Material models

Steel

The steel behaviour is described by a bilinear model with strain hardening as shown in Figure

4.5, for which the control parameters consist of the yielding strain and stress , and the

hardening or plastic modulus . Identical characteristics are assumed for both tension and

compression.

The ruling expressions both in the elastic and the hardening range are also included in Figure

4.5. However, they can be combined in a unique one, very convenient for computational pur-

poses, which can be written as follows

(4.1)

where for the elastic behaviour and for the hardening range.

Figure 4.5 Steel stress-strain relationship

zg

MyNx

xg

z

x

εx

ϕy

Acting Forces Internal Strains

yg

zg z=

y G

Cg

My ϕy,( )

Nx My εx ϕy ε ϕ

εsy fsy

Esh

fs pεs 1+/- 1 p–( )εsy+[ ]Es=

p 1= p Esh Es⁄=

fs

fsy

fsy–

εsy

εsy–

εs

EshEs

εs εsy– εsy[ , ]∈ fs⇒ Esεs=

εs εsy– εsy[ , ]∉ fs⇒ 1+/-fsy Esh εs 1+/-εsy–( )+=

Elastic range

Plastic range

1+/- sign εs( )=

εsm

fsm

Page 180: Seismic Assessment of Reinforced Concrete Frame Structures

156 Chapter 4

The plastic tensile branch is valid up to maximum stress and strain ( and ) after which

failure is assumed to occur. Usually, is taken as the uniform strain at maximum force in

tensile tests and is calculated according to the standard ENV10080 (ENV10080 (1991)).

Concrete

The model adopted to describe the concrete behaviour is schematically illustrated in Figure 4.6

and consists of the following:

• for compression, a parabolic diagram up to the peak stress, followed by a linear descending

branch until a residual stress plateau is reached;

• for tension, a linear elastic branch, after which the stress drops to zero.

Figure 4.6 Concrete stress-strain relationship

The corresponding expressions for each branch are detailed next in terms of the stress and

strain of each turning point. For the elastic tensile branch the stress is simply given by

, where is the Young modulus of concrete (initial tangent modulus or other).

For the compressive behaviour, expressions for the three distinct branches are given by

(4.2)

where refers to the slope of the linear softening branch. The defining parameters ,

and depend on the concrete strength and the peak strain obtained from compression tests

and on the degree of concrete confinement by transversal reinforcement.

fsm εsm

εsm

fc

εc

εcm

εct

εcr

fct

fcr

fcm

fc3εc( )fc1

εc( )fc2

εc( )Zm

1

fc Ecεc= Ec

fc1εc( ) fcm 2

εc

εcm-------⎝ ⎠

⎛ ⎞ εc

εcm-------⎝ ⎠

⎛ ⎞2

–=

fc2εc( ) fcm 1 Zm εc εcm–( )–[ ]=

fc3εc( ) fcr=

Zm fcm εcm

Zm

Page 181: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 157

According to Park et al. (1990), the peak stress and strain can be given by

(4.3)

in terms of the peak compressive stress and strain , resulting from cylindrical com-

pression tests. The k factor refers to the confinement degree and is expressed by

(4.4)

where stands for the volumetric confinement ratio, defined as the volume of stirrups per

unit length divided by the volume of concrete core effectively confined, and refers to the

yielding stress of transversal steel.

Following the same proposal (Park et al. (1990)), the slope can be estimated by

(4.5)

where is expressed in MPa, is the width of the confined concrete core and s is the stirrup

spacing.

For the complete definition of the softening branch, the residual stress must be also known

and is often taken as 20% of the peak stress for confined concrete.

The above introduced expressions also allow to define the diagram characteristics for uncon-

fined concrete, simply by making in Eqs. (4.4) and (4.5), along with the assumption of

a zero residual stress. Therefore, the case of unconfined concrete reduces to

(4.6)

Finally, for the tensile behaviour, the maximum tensile stress can be approximately taken

as 10% of , or according to design code prescriptions.

fcm kfc0= ; εcm kεc0

=

fc0( ) εc0

( )

k 1fsyt

fc0

------ρw+=

ρw

fsyt

Zm

Zm 0.53 0.29fc0

+145fc0

1000–-------------------------------- 3

4---ρw

b's---- kεc0

–+⎝ ⎠⎜ ⎟⎛ ⎞

⁄=

fc0b'

fcr

fcm

ρw 0=

fcm fc0= ; εcm εc0

= ; fcr 0=

Zm 0.53 0.29fc0

+145fc0

1000–-------------------------------- εc0

–⎝ ⎠⎜ ⎟⎛ ⎞

⁄=

fct( )

fc0

Page 182: Seismic Assessment of Reinforced Concrete Frame Structures

158 Chapter 4

4.2.3.3 Linear behaviour: the cracking point

For the cracking point (C) definition, it is usually considered that a RC section is at “immi-

nent” cracking when its extreme tension fibre reaches the cracking stress . Additionally, the

following behaviour assumptions are also considered: initially plane sections remain plane

after deformation (Navier-Bernoulli hypothesis); both concrete and steel behave linear elasti-

cally and all section fibres are homogenized in concrete, with the homogenizing coefficient

given by .

The study refers to the general T-section shown in Figure 4.7, of which rectangular sections are

a particular case. According to that layout, some auxiliary parameters are introduced next in

order to help computations:

• for the concrete section, and parameters are defined as and ,

respectively, such that for rectangular sections and ; the rectangular con-

crete section area is denoted by ;

• for the section steel, the following geometric moments are considered

- = total homogenized steel area;

- = total homogenized steel static moment relative to the top fibre axis ;

- = total homogenized steel moment of inertia relative to the top fibre axis .

Geometrically, the uncracked homogenized section is characterized by its total area , cen-

troid depth and total moment of inertia , given by

(4.7)

where the coefficients , and are defined as follows

(4.8)

and become equal to unity for the case of rectangular sections.

The total section axial stiffness is given by , whereas the axial stiffness for the section

without the flange is defined by .

fct

m Es Ec⁄=

Sb Sh Sb bs b⁄= Sh hs h⁄=

Sb 1= Sh 0=

bh( ) Ac

AsT

SsT

T yT( )

IsT

T yT( )

AT

dG IT

AT AcγA AsT+= ; dG

h2--- Ac

AT------γz

SsT

T

AT------+=

ITh2

3-----AcγI IsT

T ATdG2( )–+=

γA γz γI

γA 1 Sb 1–( )Sh+= γz 1 Sb 1–( )Sh2+= γI 1 Sb 1–( )Sh

3+=

EcAT( )

EcA( ) Ec Ac AsT+( )=

Page 183: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 159

Figure 4.7 Imminent cracking condition. Section, applied forces and internal strains

The moment and curvature values defining the point (Figure 4.2) are simply obtained by

applying the imminent cracking condition along with elementary static principles, provided the

external axial force is known. For that purpose, consider the section shown in Figure 4.7

with the strain diagrams corresponding to the axial force and the bending moment . The

total strain profile is given by

(4.9)

where subscripts “ ” and “ ” refer to the “axial” and “flexural” strain contributions, respec-

tively. The strain profile arises from any eccentric action of , whereas is exclu-

sively due to the action of .

In the elastic range, the concrete stress profile can be directly obtained from Eq. (4.9) and, by

imposing the tensile strength at the most tensioned fibre, the moment for the imminent

cracking condition can be given by

(4.10)

where the coordinate and the eccentricity , as defined in Figure 4.7, are given by

and , respectively; the tensile stress limit , “corrected” for the

axial force action, is defined as

(4.11)

In these conditions, the neutral axis coordinate can be also obtained from Eq. (4.9) and the

AS2

AS1

d2

d1h y

z

Cg

GdGdg

zt

eN

zg

xg

NM

zzz

xN

MNMze

de

εfNεf

M εaN

neutral axis

Partial strain diagramsSection Applied forces

hs

bs

b

ASj

I

djI

yT

C

N

N M

ε z( ) εaN εf

N z( ) εfM z( )+ +=

a f

εfN z( ) N εf

M z( )

M

fct Mc

Mc fct'IT

zt---- NeN–=

zt eN

zt h dG–= eN dg dG–= fct'

fct' fctNAT------–=

ze

Page 184: Seismic Assessment of Reinforced Concrete Frame Structures

160 Chapter 4

corresponding depth (referring to the most compressed fibre) becomes

(4.12)

The cracking curvature can be obtained by

(4.13)

and the uncracked flexural stiffness becomes defined as

(4.14)

which represents the slope of the M-ϕ curve branch up to the cracking point C.

From Eq. (4.14) it is apparent that the axial force may affect the flexural stiffness due to

any eccentricity; therefore, unless the section is symmetric or the axial force is null, different

flexural stiffness is obtained for positive and negative bending directions.

The expressions obtained for , and refer to the positive direction of bending. For the

negative one, the same expressions are still valid if is replaced by , but still refers

to the top fibre which is now the most tensioned one.

4.2.3.4 Turning points for non-linear behaviour

Basics for a unified procedure

By contrast with the closed form expressions obtained for the linear behaviour, the determina-

tion of any point for the post-cracking curve part requires a more complex and elaborate proce-

dure. However, for any point to be found (yielding, ultimate or others), a unified procedure can

be adopted based on the following:

• A pre-defined criterion, strictly associated to each turning point, states a given strain for

a specified fibre located at depth (measured from the most compressed fibre) as shown

in Figure 4.8; this provides a fixed point around which the strain linear profile can twist,

thus only another point remaining to be determined, which usually corresponds to the neu-

tral axis fibre at depth .

de

de dGN

fct'AT------------zt⎝ ⎠

⎛ ⎞–=

ϕc( )

ϕcεct

h de–--------------=

EcIT( )∗Mc

ϕc------- EcIT

1 NeN

Mc---------+⎝ ⎠

⎛ ⎞-------------------------= =

EcIT

Mc ϕc de

zt zt h– de

εa

da

d0

Page 185: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 161

• The equilibrium equations (for axial force and bending moment) are written for the previ-

ous paragraph conditions, but, since the material (steel and concrete) stresses depend piece-

wise non-linearly on the strain profile, one trial value of must be used to define the

adequate stress profile for each distinct fibre zones.

Actually, this means that a set of assumptions has to be made, concerning the behaviour of

characteristic zones of the section, which still must be verified for the solution of the equi-

librium equations.

Therefore, with the initial assumptions set, the equilibrium equations can be written and

solved for a new value of ; this, in turn, leads to a new set of section behaviour conditions

which must match the assumptions made a priori; iterations are performed until this coinci-

dence occurs.

• Once the value for equilibrium is found, the strain profile is defined and both curvature

and bending moment can be computed.

Using such a scheme, the only aspect that modifies from one point (e.g yielding) to another

(e.g. ultimate) is the criterion stating the specific conditions for each of them.

For the following developments, consider again the generic T-shape section shown in Figure

4.8, where a given strain is imposed at depth . For non-linear behaviour, the section is

considered fully-cracked, thus with the concrete below the neutral axis assumed inactive.

Both the applied forces ( and ) and internal ones developed in each steel layer ( ,

and ) and in the concrete active (shaded) area ( ) are indicated in Figure 4.8.

Figure 4.8 Generic situation for the unified procedure. Forces and internal strains

For these conditions, the curvature, the strain at any fibre (of generic coordinate ) and the

d0

d0

d0

εa da

N M Fs1Fs2

Fsj

I Fc

AS1

d2

d1hy

z

Cg

O

d0dg

zg

xg

NM

zz

x

da

ε z( )

εa

neutral axis

Strain diagramSection Applied forces

AS2

FS2

FS1

Fc

ASj

I

djI

FSj

I

Internal forces

dc

z

Page 186: Seismic Assessment of Reinforced Concrete Frame Structures

162 Chapter 4

equilibrium equations can be written in terms of the neutral axis depth as the main variable,

and are given by

(4.15)

(4.16)

(4.17)

(4.18)

where stands for the moment of each force, taken relative to the neutral axis, and is

the number of the secondary steel reinforcement layers. Each contribution from concrete and

steel for Eqs. (4.17) and (4.18) is detailed in the next sections.

Concrete contribution

Unconfined and confined zones

The concrete active area usually consists of unconfined and confined zones which have to be

considered separately due to their different behaviour models.

For both the T-shape and the rectangular sections, the geometry of that active area, either the

unconfined or the confined zone, can be always reduced to a T-shape form. This means that

contributions from both zones can be handled in the same manner, if adjustable dimensions are

adopted (flange width , web width , top fibre depth and flange thickness ) which are

assigned the adequate values to match the desired configuration.

Figures 4.9 and 4.10 show the concrete section division into unconfined and confined zones,

for T-shape and rectangular sections. Their reduction to convenient T-shape sections and the

corresponding geometric characteristics ( , , and ) are also included, with the super-

scripts “u” and “c” standing for the unconfined and confined parts, respectively.

It should be noted that in the T-shape section type (see Figure 4.10) the two lateral dashed areas

d0

ϕ ϕ d0( )εa

da d0–----------------= =

ε z( )εa

da d0–----------------z=

N Fc F+ s1Fs2

Fsj

I

j 1 nI,=∑+ +=

N dg d0–( ) M+ M Fc( ) M Fs1( ) M Fs2

( ) M Fsj

I( )j 1 nI,=∑+ + +=

M …( ) nI

bs bw dt hs

bs bw dt hs

Page 187: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 163

are considered in the confined part, even if they are not enclosed by stirrups, since the presence

of the slab provides a certain confinement to the lateral zones.

Figure 4.9 Rectangular section split into unconfined and confined zones

Figure 4.10 T-shape section split into unconfined and confined zones

Formulation for a T-shape section contribution

The definition of a general formulation for the concrete contribution to the equilibrium equa-

tions is not an easy task due to the non-linear piecewise stress distribution and to the specific

features of the section geometry.

However, for a certain type of sections, it may be possible the statement of a unique formula-

tion able to handle all the distinct situations likely to appear. In order to render easier the iden-

c2

h

d0

Section

c1

b

cl cl 2cl

b b 2cl–

Unconfined Confined

dtu

dtc Unconfined

Confined

bsc b 2cl–= bw

c bsc=

hsc 0= dt

c c2=

bsu b= bw

u 2cl=

hsu c2= dt

u 0=n.a.

c2

h

d0

Section

c1

b

cl cl 2cl

bs

b 2cl–

Unconfined Confined

dtu

dtc

bsc bs 2cl–= bw

c b 2cl–=

hsc 0= dt

c c2=

bsu bs= bw

u 2cl=

hsu c2= dt

u 0=

n.a.

bs bs 2cl–

Page 188: Seismic Assessment of Reinforced Concrete Frame Structures

164 Chapter 4

tification of these situations, the influence of stress non-linearity and of geometry peculiarities

must be separated as much as possible; this can be accomplished by the following procedure:

• The stress distribution over the active height is divided in zones where the stress function is

uniquely defined (herein termed by “stress zones”); the development of this distribution

and, consequently, the number of “stress zones” depend on the maximum compressive

strain which, in turn, depends on the trial value as shown in Figure 4.11. Actually, the

depths and , where the deformations and can be found, are readily obtained

through Eq. (4.16) and given by

(4.19)

Figure 4.11 Influence of d0 in the development of stress zones

• The active area is also divided into zones where the geometry is uniquely defined (called

“geometric zones”); the number of these zones also depend on the value, since it influ-

ences the active area as shown in Figure 4.12 for the general T-shape section.

Figure 4.12 Influence of d0 in the development of uniform geometry zones

d0

dm dr εm εr

dm d0 zm+= with zmεm

εa----- da d0–( )=

dr d0 zr+= with zrεr

εa---- da d0–( )=

d0

dt

n.a.

d0

d0

n.a.

n.a.

zm

zt

zr fcm

fcm

fcr

z fc

fc

fczm

z z

13 2

2 1

1

dr

dm

a) 3 stress zones b) 2 stress zones c) 1 stress zone

d0

dt

d0

d0

n.a.

n.a.ds

d0 ds> 2 zones⇒ d0 ds≤ 1 zone⇒

Page 189: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 165

• The intersection of the “stress zone” and “geometric zone” sets, leads to a final set of zones

where the force and moment resultants can be obtained by analytical integration (the “inte-

gration zones”). For the rectangular section, due to the geometric uniformity all over the

section, the “integration zones” coincide with the “stress zones”. In the case of the T-shape

section there is only one change in the geometric uniformity due to the flange-web transi-

tion at depth (see Figure 4.12). Therefore, there is at most one more “integration zone”

than “stress zones”, whose position depends on the trial value of as shown in Figure

4.13.

Figure 4.13 Development of integration zones and adjustable widths bi and bi

• The fact that only one change in geometry occurs for T-shape sections greatly simplifies the

integration procedure. For each “stress zone” two width values are set up, namely for the

less compressed (inferior) and for the more compressed (superior) fibres, respectively

denoted by and , where the subscript i stands for the zone number. If these widths

ds

d0

dt

ds

d0

n.a.zm

zt

zr

z

fc

2

drdm

d0

n.a.zm

zt

zr

z

fc1

3

drdm

2

dt

ds d0

n.a.zm

ztzr

z

fc

drdm d0n.a.

zm

ztzr

z

fc

1

drdm

b

bs

Section

b3 bs=

b2 b=

b1 b=

b3 bs b–( )–=

b2 0=

b1 0=

b3 bs=

b2 bs=

b1 b=

b3 0=

b2 bs b–( )–=

b1 0=

b3 bs=

b2 bs=

b1 bs=

b3 0=

b2 0=

b1 bs b–( )–=

b3 bs=

b2 bs=

b1 bs=

b3 0=

b2 0=

b1 0=

3

1

3

1

2

2

12

33

00

0

0

zs zs

zs

biinf bi

sup

Page 190: Seismic Assessment of Reinforced Concrete Frame Structures

166 Chapter 4

are equal, the stress integration is straightforward because no geometry change occurs in

that zone. By contrast, if these widths are different, it means that the geometry transition

fibre exists in that zone; therefore, a first integration is performed over the whole zone

height for a constant width , which is then corrected with the result from a fur-

ther integration performed between the inferior and the transition fibres, for a constant

width . Figure 4.13 includes different situations and the corresponding

widths and for each integration zone, in order to clarify this procedure.

The numerical implementation of this algorithm requires that, for each zone, some integrals be

previously obtained analytically, viz:

• the total force and moment per unit width, resulting from integration between the

inferior and the superior fibres and given by

(4.20)

• the correcting force and moment per unit width, obtained by integration between

the inferior fibre and the transition one (that may exist in the zone), and written as

(4.21)

Eq. (4.16) is used to replace in the integrals of Eqs. (4.20) and (4.21), which allows to

express them in terms of their limits and the main variable . The analytical integration is

trivial since only polynomials are involved, but the results are not included here because cum-

bersome expressions are obtained.

Therefore, once a trial value is assumed, the corresponding and depths can be calcu-

lated by Eq. (4.19) and the integration zones can be defined accordingly. It is checked which of

bi bisup=[ ]

bi bisup bi

inf–( )–=[ ]

bi bi

fi( ) mi( )

f1 f1 d0( ) fc1εc( ) zd

zm

0

∫= = ; m1 m1 d0( ) fc1εc( )z zd

zm

0

∫= =

f2 f2 d0( ) fc2εc( ) zd

zr

zm

∫= = ; m2 m2 d0( ) fc2εc( )z zd

zr

zm

∫= =

f3 f3 d0( ) fc3εc( ) zd

zt

zr

∫= = ; m3 m3 d0( ) fc3εc( )z zd

zt

zr

∫= =

fi( ) mi( )

f1 f1 d0( ) fc1εc( ) zd

zs

0

∫= = ; m1 m1 d0( ) fc1εc( )z zd

zs

0

∫= =

f2 f2 d0( ) fc2εc( ) zd

zs

zm

∫= = ; m2 m2 d0( ) fc2εc( )z zd

zs

zm

∫= =

f3 f3 d0( ) fc3εc( ) zd

zs

zr

∫= = ; m3 m3 d0( ) fc3εc( )z zd

zs

zr

∫= =

εc

d0

d0 dm dr

Page 191: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 167

the four situations of Figure 4.13 is matched for the trial case values, by searching the stress

zone where the transition fibre is lying; the result of this search defines the assumption related

with the concrete behaviour, as already referred.

Note that, in case one or both depths and happen to lie above the top fibre depth ,

the corresponding zones (2 or 3) fall outside the section geometry. Therefore, their contribu-

tions must not be considered for equilibrium, which is achieved by limiting such depths to the

minimum value ; in these conditions, the result of the involved stress integrals vanishes

automatically. However, from the computational standpoint, it is more convenient that Eqs.

(4.20) and (4.21) yield expressions always in terms of the same variables and, consequently,

the following alternative expressions for the integration limits are more appropriate:

(4.22)

(4.23)

where, according to Eq. (4.19), and are still given by

(4.24)

Replacing the integration limits and by and , respectively, it is ensured that inte-

grals are expressed only in terms of (as main variable) and the parameters and .

For each zone, the total and correcting widths, respectively, and , are computed in order

to affect the corresponding contributions of force and moment integrals (per unit width) to the

concrete resultants and (see Eqs. (4.17) and (4.18)). Finally, these resultants can be

generally expressed, as functions of , by

(4.25)

(4.26)

where, only the integrals associated with non-zero width need actually to be evaluated.

dm dr dt( )

dt

zm* ξmdm 1 ξm–( )dt d0–+= with

ξm 1= if dm dt>

ξm 0= if dm dt≤⎩⎨⎧

zr* ξrdr 1 ξr–( )dt d0–+= with

ξr 1= if dr dt>

ξr 0= if dr dt≤⎩⎨⎧

dm dr

dm d0εm

εa----- da d0–( )+= and dr d0

εr

εa---- da d0–( )+=

zm zr zm* zr

*

d0 ξm ξr

bi bi

Fc M Fc( )

d0

Fc Fc d0( ) bifi bifi+( )i=1,3∑= =

M Fc( ) Mc d0( ) bimi bimi+( )i=1,3∑= =

Page 192: Seismic Assessment of Reinforced Concrete Frame Structures

168 Chapter 4

Unconfined and confined concrete contributions

Using the formulation and the division of zones as described above, the contribution of both

unconfined and confined parts is straightforwardly defined, since they have been reduced to

general T-shape configurations.

The geometric and material model parameters of each part just have to be introduced in the

expressions presented above and the respective contributions to and become readily

defined. Some specific conditions may have to be considered for the unconfined or the con-

fined parts, depending on the criterion assumptions and requirements for each point definition.

The global contribution of concrete to the equilibrium Eqs. (4.17) and (4.18) is finally obtained

by the superposition of the unconfined and confined concrete contributions, and given by

(4.27)

(4.28)

which are rational polynomials of . Again, the superscripts “u” and “c” refer to the uncon-

fined and confined parts, respectively.

Steel contribution

Several steel layers are considered: the principal ones (top and bottom reinforcement, to which

some point definition criteria usually refer) and the interior layers that may exist, as shown in

Figure 4.3.

The contribution of each generic layer (principal and secondary ones included) is defined

according to the steel stress-strain diagram shown in Figure 4.5 and traduced by Eq. (4.1). The

behaviour can be either in the linear elastic or in the positive or negative plastic ranges, which

is checked-out according to the trial value of .

From Eq. (4.16) the strain , at each layer at depth , is given by

(4.29)

Fc M Fc( )

Fc Fc d0( ) Fcu Fc

c+ biufi

u biufi

u+( ) bicfi

c bicfi

c+( )+[ ]i=1,3∑= = =

Mc d0( ) M Fcu( ) M Fc

c( )+ biumi

u biumi

u+( ) bicmi

c bicmi

c+( )+[ ]i=1,3∑= =

d0

d0

εsii di

εsi

εa

da d0–---------------- di d0–( )=

Page 193: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 169

and the decision conditions are the following

(4.30)

The assumption concerning the steel behaviour becomes defined by the values of and .

Subsequently the stress in the generic steel layer can be expressed in terms of , and,

finally, the force and moment contributions to the equilibrium equations are given by

(4.31)

(4.32)

where is the corresponding steel area.

Solution of the equilibrium equation and corresponding section state

Using the contributions of concrete and steel as above described, Eq. (4.17) can be expressed

in terms of . A detailed analysis of each contribution shows that it consists of a rational pol-

ynomial of , which can be reduced to a polynomial form of 3rd order if all the equation is

multiplied by a factor .

Thus, the steel and concrete (unconfined and confined) contributions for the global equation

are factored by this term and the corresponding polynomial partial coefficients are obtained

accordingly. The same applies for the axial force term which leads to a 2nd order polynomial.

The final 3rd order equation is solved by the Cardan algorithm (Spiegel (1970)), thus leading to

1, 2 or 3 distinct real roots, depending on the equation coefficients.

However, the obtained roots are not necessarily solutions of the equilibrium problem. Actually,

a calculated value of will be the desired solution if it fulfils the following requirements:

• to be an admissible solution, meaning that it has to lie within a range of values whose

upper and lower bounds are pre-defined according to the criterion for the

point definition;

• to be an equilibrium solution, for which the equilibrium equation assumptions based on the

εsiε– sy εsy[ , ]∈ => pi 1= (Elastic)

εsiε– sy εsy[ , ]∉ =>

pi p=

1i+/- sign εsi

( )=⎩⎨⎧

(Plastic)

pi 1i+/-

fsid0

FsiFsi

d0( ) pi εadi d0–da d0–----------------⎝ ⎠

⎛ ⎞ 1i+/- 1 pi–( )εsy+ EsAsi

= =

M Fsi( ) Msi

d0( ) Fsidi d0–( )= =

Asi

d0

d0

da d0–( )2

d0

d0

d0max( ) d0

min( )

Page 194: Seismic Assessment of Reinforced Concrete Frame Structures

170 Chapter 4

trial value, still must be verified for the obtained admissible solution.

The solution searching technique can be either a “sequential” one where the whole admissible

solution range is checked up step-by-step until the equilibrium solution is found,

or an “oriented” one where new trial values are based on the equilibrium equation residuals of

the previous obtained roots. In the present work the solution is first searched by the “oriented”

technique and, only in case that it is not found after a given maximum number of iterations, the

“sequential” technique is activated; details of both techniques can be found in Arêde and Pinto

(1996).

Once the equilibrium solution is found for the neutral axis depth (say ), the section state can

be readily known. The installed curvature is obtained by Eq. (4.15) for ; the corre-

sponding moment is obtained from Eq. (4.18), where all the concrete and steel contribu-

tions are given by Eqs. (4.28) and (4.32), respectively, after being particularized for .

The deformation of any concrete or steel layer is given by Eq. (4.16) (for ) and the corre-

sponding stress is obtained according to the respective model. In particular, for the steel layers,

the force resultant can be readily computed using Eq. (4.31).

Note that all the development has been made referring to the positive bending direction. For

the negative one, the same procedure can be applied, having in mind that the negative bending

in a given section is equivalent to the positive one in the inverted section. Therefore, the fol-

lowing adjustments are due before the negative bending calculations start:

• for T-shape sections only:

is replaced by and vice-versa; is replaced by ;

• for all types of sections:

are replaced by ; are replaced by ;

are replaced by .

In such conditions, the obtained moment and curvature are positive quantities and, therefore,

the sign must be changed “a posteriori”. However, it is reminded that the value (positive)

still refers to the most compressed fibre (in this case, the bottom one) and, thus, any computed

layer deformations and forces will come already with the correct sign (positive for tension and

negative for compression).

d0

d0min d0

max,[ ]

d0*

ϕ* d0 d0*=

M*

d0*

d0*

b bs hs h hs–

As1d1,( ) As2

h d– 2,( ) As2d2,( ) As1

h d– 1,( )

djI h dj

I–

d0*

Page 195: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 171

Criteria for definition of turning points

The scope of the following paragraphs is the definition of the physical conditions associated

with the characteristic points of the diagram to be determined. These points are usu-

ally the yielding and the ultimate ones, and are related to the attainment of yielding or ultimate

conditions in one of the constitutive materials. In the present study, these conditions are con-

sidered as in Pipa (1993) and are detailed next. Besides these main points, supplementary ones

can be also defined aiming at a better knowledge of the post-yielding behaviour.

For every point, a given strain is prescribed for a specific layer and a range of possible neutral

axis positions is set up to restrict solutions to the physically acceptable ones. It is worth men-

tioning that, for the present stage of algorithm development, only common situations are con-

sidered, in the sense that neither totally tensioned nor completely compressed sections are

admitted. This restriction is quite acceptable because trilinear envelopes are usually obtained

for the structure with vertical static loads, where the predominance of tensile or compressive

axial force at yielding or ultimate stages is undesirable from the seismic design point of view;

indeed, under axial tensile forces the flexural strength can be drastically reduced, while under

dominant compressive force both the strength and the ductility may be seriously affected.

Yielding point

The yielding point is defined according to one of the two following criteria:

• the most tensioned steel layer reaches the yielding strain

yielding of steel;

• the most compressed layer of confined concrete reaches the peak strain

yielding of concrete, where the superscript “c” stands for the confined model.

The first criterion generally holds for beam sections, where the axial force is null, or very low,

leading to a low level of the maximum compressive strain in the concrete. By contrast, the sec-

ond criterion applies for column sections, where the axial force may induce considerably high

levels of compressive strain; in such cases, when this criterion condition is matched, the tensile

strain in the steel is usually quite far from the yielding strain.

Figure 4.14 illustrates generic strain profiles associated to each yielding criterion; the fixed

point, in turn of which the strain diagram can twist, is indicated, as well as the boundary

M ϕ–( )

εsy

εcmc

Page 196: Seismic Assessment of Reinforced Concrete Frame Structures

172 Chapter 4

dashed lines for the possible strain diagrams which provide the neutral axis upper and lower

bounds ( and ).

For the first criterion, the strain is imposed at the layer depth . The maxi-

mum value of is reached when the strain at the most compressed confined concrete layer is

equal to and the corresponding strain diagram defines the transition between the two crite-

ria. Therefore, can be obtained through Eq. (4.16) and is given by

(4.33)

The minimum depth is set up at the top fibre , since it corresponds to a limit for the

acceptable solutions. For positions of neutral axis above it, all the fibres are tensioned which

corresponds to an undesirable situation of dominant tensile axial force as pointed out above.

Figure 4.14 Yielding criteria at the tensioned steel or at compressed confined concrete

Instead, for the second criterion the strain is imposed at the layer depth .

The value of corresponds now to the of the previous criterion (given by Eq. (4.33)),

while the maximum depth is now set up at the bottom fibre . Beyond this limit, all

the fibres are compressed at yielding condition, which means predominance of compressive

axial force, i.e. another undesirable solution already referred.

Both criteria are analysed using the unified procedure explained before, and the yielding point

is associated with the criterion leading to the least curvature.

d0max d0

min

εa εsy= da d1=

d0

εcmc

d0max

d0max d1εcm

c dtcεsy–

εcmc εsy–

-------------------------------=

d0min 0=( )

c2

h

d0

Section

c1

dtu

dtc

n.a.

z

da

ε z( )

εa εsy=

d0min

d0max

εcmc

d1

d0

n.a.

z

daε z( )

εsy

d0min

d0max

εa εcmc=

b) Yielding at concretea) Yielding at steel

Twisting point

Confinedconcrete

εa εcmc= da dt

c=

d0min d0

max

d0max h=

Page 197: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 173

Ultimate point

The ultimate point criteria are defined in a similar way as for the yielding point, and can be

stated as follows:

• the most tensioned steel layer reaches a pre-defined maximum strain

ultimate at steel;

• the most compressed confined concrete layer reaches the residual strain

ultimate at concrete.

The maximum steel strain is defined according to 4.2.3.2, whereas is the strain when

the residual stress is first reached in the confined concrete model (see Figure 4.6) and can be

obtained imposing in Eq. (4.2).

Again, the first criterion generally holds for beam sections, while the second applies for col-

umn sections where the axial force can be important; both are shown in Figure 4.15.

Figure 4.15 Ultimate criteria at the tensioned steel or at compressed confined concrete

For the first criterion, the strain is imposed at the layer depth , and the upper

bound is still given by Eq. (4.33), but with and replaced by and , respec-

tively; the lower bound, is still set up at the top section fibre as for the yielding point.

For the second criterion the strain is imposed at the layer depth ; again the

value of corresponds to the of the previous criterion and the maximum depth is still

set up at the bottom fibre , due to the same reasons explained for the yielding point.

However, special care is taken with the unconfined part in order to eliminate the corresponding

εsm

εcrc

εsm εcrc

fc2

c εcrc( ) fcr

c=

c2

h

d0

Section

c1

dtu

dtc

n.a.

z

da

ε z( )

εa εsm=

d0min

d0max

εcrc

d1

d0

n.a.

z

daε z( )

εsm

d0min

d0max

εa εcrc=

b) Ultimate at concretea) Ultimate at steel

Twisting point

Confinedconcrete

εa εsm= da d1=

d0max εsy εcm

c εsm εcrc

d0min

εa εcrc= da dt

c=

d0min d0

max

d0max h=

Page 198: Seismic Assessment of Reinforced Concrete Frame Structures

174 Chapter 4

flange, which, at this deformation stage, is already spalled off.

As for the yielding point, both criteria are checked and the ultimate point is associated with the

least curvature.

Supplementary point

Although for some sections (e.g., beam ones) the post-yielding monotonic behaviour is quite

well represented by the straight line between the yielding (Y) and ultimate (U) points, for oth-

ers a more detailed knowledge is required in the post-yielding zone. This can be accomplished

by applying the unified procedure to define supplementary points between Y and U, provided

the physical conditions for them are appropriately defined.

In the following, an example of a supplementary point is included, whose definition criterion

states that the strain at the most tensioned steel layer is the average of strains corresponding to

the Y and U points. Denoting these strains by and , the imposed strain at is

given by and defines the twisting point of the strain diagram, as shown in

Figure 4.16.

Figure 4.16 Supplementary point. Criterion related to the most tensioned steel

The upper and lower bounds for are set up imposing that the strain at the most compressed

layer of confined concrete must be within the values obtained for the yielding and ultimate

conditions (say and , respectively); therefore, according to the dashed line strain dia-

grams of Figure 4.16, and are now given by

εsY εs

U da d1=

εa εsY εs

U+( ) 2⁄=

c2

h

d0

Section

c1

dtu

dtc

n.a.

z

da

ε z( )

εa εsY εs

U+( ) 2⁄=

d0min

d0max

εcU

d1

εcY

Twisting point εsY

εsUConfined

concrete

d0

εcY εc

U

d0min d0

max

Page 199: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 175

(4.34)

Note that the same reasoning can be used for any point between Y and U, and, in the limit, the

post-yielding zone can be completely traced out by defining a set of supplementary points

whose imposed strain at the tension steel layer is , where ranges

between 0 and 1.

The criterion can be also related to an imposed strain at the most compressed concrete layer,

following an entirely similar strategy; in such case, and are associated with the ten-

sion steel strains for the Y and U points.

4.2.3.5 Remarks on implementation and validation

The above described algorithm has been implemented in CASTEM2000 by means of a new

operator named TRIL, supported by a specific new driver transforming input/output data struc-

tures and controlling a master subroutine TRILIN where calculations are performed.

Basically, the input consists of the material model data, the geometric characteristics, the lon-

gitudinal and transversal steel data and the axial force. The output provides the initial elastic

characteristics (axial and flexural stiffness), the M-ϕ curve defined by its turning points and

further extra results related with the internal section state for each turning point. Details on

implementation can be found in Arêde and Pinto (1996).

The algorithm validation has been carried out by analysing several examples of reinforced con-

crete sections, for which the trilinear approximations are compared against the results of the

section monotonic analysis using fibre model discretization with the same material model data.

Several section characteristics, related with reinforcement contents, slab participation and axial

force, are varied in order to assess the trilinear approximation throughout some typical cases of

sections. Results of such validation are reported in Arêde and Pinto (1996), showing that the so

obtained trilinear curves compare quite well with the M-ϕ curves obtained by fibre analysis.

d0min dt

cεa daεcY–

εa εcY–

--------------------------= and d0max dt

cεa daεcU–

εa εcU–

---------------------------=

εa λεsY 1 λ–( )εs

U+= λ

d0min d0

max

Page 200: Seismic Assessment of Reinforced Concrete Frame Structures

176 Chapter 4

4.3 Flexibility-based element validation at the single member level

4.3.1 General

The exhaustive validation of a global element model such as the flexibility based model devel-

oped herein, is a rather difficult and cumbersome task to perform, particularly if comparison

against experimental evidence is sought. The wide range of situations that can be considered in

one element integrated in a complete structure can hardly be reproduced and found in experi-

ments on single members or sub-assemblages where the interpretation of phenomena is often

easier.

Nevertheless, the numerical simulation of single member experiments by the flexibility ele-

ment model is quite useful because the main features of the model response can be assessed by

comparison against experimental results and any model limitations can be detected.

The present section deals with the numerical simulation of several reinforced concrete mem-

bers experimentally tested under monotonic and cyclic quasi-static loading conditions,

included in the preliminary experimental campaign of small scale tests for supporting the full-

scale tests of the four storey building performed at the ELSA laboratory (Negro et al. (1994)).

This member testing aimed at representing the beam behaviour in the full-scale structure, and

particularly, the assessment of slab participation was sought. Thus, several cantilever beams

having identical section reinforcement contents were tested, both with and without a slab

flange contributing to the beam behaviour. Tests for T-shape sections were carried-out by Pipa

and Carvalho (Pipa (1993), Carvalho (1993)) at the “Laboratório Nacional de Engenharia

Civil” (LNEC), Lisbon, while those for rectangular sections were performed by König and

Heunish (Carvalho (1993)); a full description of specimen layouts, material characteristics,

testing setup and results can be found elsewhere (Carvalho (1993)).

4.3.2 Specimen characteristics and test description

Two basic section configurations were considered, depending on the reinforcement layout: the

section type S1, assumed representative of external frame beams, and the section type S2 for

the internal frame beams, both at the first storey of the referred building where beams in the

test direction were more heavily reinforced.

Page 201: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 177

In the present work, only the section type S2 members are analysed, with both the rectangular

and the T-shape configurations. The respective section layout and the schematic structural rep-

resentation of members are shown in Figure 4.17; T-shape section beams are referenced as

LNEC-beams and rectangular ones are designated KH-beams.

Figure 4.17 Section layout S2 and schematic representation of tested cantilever beams

Specimens were designed and built-up at a 2/3 reduced scale and, for both beam configura-

tions, material characteristics were taken similar to those specified for the full-scale structure.

Reinforcing bars were B500S Tempcore steel, both in the longitudinal and transversal and slab

reinforcement, for which standard tension tests were made in order to evaluate the mechanical

characteristics. Concrete was taken of the class C25/30, using the same mix in LNEC and KH

beams for conformity of mechanical characteristics, which were evaluated by standard com-

pressive strength tests on 150 mm cubes for each cast operation.

From the available data (Pipa (1993) and Carvalho (1993)), average characteristics were

extracted and adopted in the numerical simulations as summarized in Table 4.1 (the same nota-

tion is used as in 4.2.3.2).

Concerning the steel, and stand for the tensile strength and to the strain at maximum

force, respectively; identical characteristics were used for longitudinal and transversal steel

due to their similarity.

0.10

1.00

a) LNEC beam section S2 b) KH beam section S2

0.02

0.03

0.20

F+

c) Cantilever

4φ12

2φ12

4φ6 4φ6

0.30y

z

(x)M+

0.02

0.03

0.20

4φ12

2φ12

0.30y

z

(x)M+

yz

x

1.50

0.15 0.15

M+ϕ+

lp1lp2

δ+

φ6 // 0.07

φ6 // 0.07

beam

fsm εsm

Page 202: Seismic Assessment of Reinforced Concrete Frame Structures

178 Chapter 4

For the concrete, both the Young modulus and the tensile strength were estimated from

the characteristic strength , by means of code expressions as used in 6.2.2 for the full scale

structure. Additionally, for KH beams, the values indicated with (*) were taken equal to those

for LNEC beams due to lack of test information. It is noteworthy that, with exception of , all

listed data refer to mean values obtained from the various material samples.

Tests have consisted of monotonic ones (in both loading directions) and of quasi-static cyclic

sequences of imposed displacements as follows:

• Monotonic tests, referenced as V1 and V2, respectively for the positive and negative direc-

tions (thus, inducing tensile and compressive strains in the bottom reinforcement, respec-

tively), were performed aiming at the assessment of ultimate displacements ( and ).

• Four cyclic tests, viz V3, V4, V5 and V6, were done by applying tip displacement

sequences whose peak values were normalized by the above referred ultimate displace-

ments. These loading sequences were defined for the purpose of evaluating the influence of

the imposed displacement level and the cyclic repetition effect on the specimen behaviour.

In the present work, both monotonic tests were numerically simulated, while only the V5 and

V6 cyclic ones were analysed because they were considered representative for the purpose of

cyclic tests. According to Carvalho (1993), slightly different loading sequences were used for

the two types of beams, which can be confirmed by inspection of Figure 4.18 where the cyclic

histories of imposed displacements are depicted in terms of the monotonic ultimate displace-

ments and (also included in the figure).

The numerical simulations were not performed over each complete sequence. Indeed, for very

large imposed displacements, buckling of reinforcement bars was triggered-off, which cannot

be simulated by the numerical model; therefore, from that stage on, the test development is

meaningless from the numerical simulation standpoint and it has not been included here (the

respective breakpoint is also indicated in Figure 4.18).

Table 4.1 Mechanical properties of steel and concrete of LNEC and KH beams

Steel Concrete

(MPa) (%) (MPa) (%) (%) (MPa) (%) (MPa) (MPa) (GPa)LNEC 550 0.27 615 9.8 0.34 35.0 0.2 27.8 31.3 2.8

KH 538 0.27 632 10.7 0.45 35.8 (*)0.2 (*)27.8 (*)31.3 (*)2.8

Ec fct

fck

fck

fsy εsy fsm εsm Esh Es⁄ fc0εc0

fck Ec fct

δu+ δu

-

δu+ δu

-

Page 203: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 179

Note that, since for each test a different specimen was used, slight variations of material prop-

erties are obviously expectable; therefore, this has to be kept in mind when comparing experi-

mental results with numerical simulations, actually performed with identical average

characteristics for all tests.

Figure 4.18 Cyclic sequences (V5 and V6) of imposed displacements for LNEC and KH beams

Tests have provided results for the applied force F (measured by means of force transducers)

and for the imposed displacement , as well as for the upper and lower beam face displace-

ments in the potential plastic zone lengths, indicated in Figure 4.17-c) by and , which

were obtained by displacement transducers. The plastic zone lengths refer to , where is

the beam depth, in correspondence with the often accepted plastic zone spreading between

and .

From these measurements, tip force-displacement curves and moment - average curvature dia-

grams in the plastic zones were obtained and are used herein for comparison with numerical

results.

a) LNEC beams

b) KH beamsδ δu+⁄

δ δu-⁄

δ δu-⁄

δ δu+⁄

Cycle Cycle

CycleCycle

0.8

0.6

0.4

0.2

0.0

0.8

0.6

0.4

0.2

0.8

0.6

0.4

0.2

0.0

0.8

0.6

0.4

0.2

0.8

0.6

0.4

0.2

0.0

0.8

0.6

0.4

0.2

0.8

0.6

0.4

0.2

0.0

0.8

0.6

0.4

0.2

δu- 154mm–=

δu+ 107mm=

δu- 132mm–=

δu+ 109mm=

V5

V5

V6

V6

Not simulated Not simulated

Not simulated Not simulated

δ

lp1lp2

h 2⁄ h

h 2⁄ h

Page 204: Seismic Assessment of Reinforced Concrete Frame Structures

180 Chapter 4

4.3.3 Numerical simulations

For the numerical analyses subsequently presented, the material characteristics listed in Table

4.1 were adopted for use with the steel and concrete models described in 4.2.3.2. Additionally,

the following assumptions were made:

• uniform concrete cover of 2 cm and volumetric confinement ratio of 0.8%, obtained accord-

ing to the stirrup contents shown in Figures 4.17-a) and b);

• the slab reinforcement of LNEC beams is included in the top steel layer area, having the

same properties of the main reinforcement;

• centroids of both the top and the bottom reinforcement layers are located at 3 cm from the

nearest section face.

The moment-curvature skeleton curve for each section was obtained by the TRILIN operator

presented in to 4.2.3 and, for numerical simulation using the flexibility element model, a single

element was considered in correspondence with the schematic representation shown in Figure

4.17-c) and with uniform section characteristics.

The availability of experimental measurements of average curvature in the plastic zones has

suggested a preliminary comparison of the experimental moment-curvature relationships with

the numerical trilinear approximation ones as obtained by the TRILIN operator. Additionally,

this comparison was complemented with fibre analysis of the section, performed by means of

the same numerical tool as used for the section analysis included in 3.5.1.

The corresponding monotonic moment-curvature diagrams (for both loading directions) are

depicted in Figure 4.19 and compared against experimental curves obtained from V1 and V2

tests, for the plastic zone referenced by in Figure 4.17-c). The fibre (indeed, layer) discreti-

zations of sections are also included in Figure 4.19.

From these result comparisons, the following aspects are highlighted:

• There is good agreement between the trilinear approximation and the fibre analysis curves,

for the stages where such agreement is sought (viz for the turning points); naturally, outside

such stages, deviations do occur as is apparent in the immediate post-yielding range where

the hardening effect cannot be followed by that step-wise linear curve, unless a supplemen-

tary branch is considered.

lp1

Page 205: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 181

Figure 4.19 Comparison of numerical and experimental monotonic M-ϕ curves for LNEC and

KH beam sections

• The approximation for yielding moments is quite acceptable; indeed, although it may

appear somewhat crude for the negative bending of the LNEC beam, it actually fits within

the scatter of the steel yielding stress obtained from material tests (a maximum deviation of

8% from the average value of can be found in the tested sample set (Carvalho (1993))).

• Instead, yielding curvatures are significantly underestimated by the numerical analyses,

which is due to the pull out effect caused by reinforcement slippage inside the anchorage

zone and included in the deformations measured in the first plastic zone . The effect is

extensively reported in Pipa (1993) as responsible for yielding displacements about 50%

.0 .10 .20 .30 .40 .0

10.0

20.0

30.0

40.0

50.0

.0 .10 .20 .30 .40 .0

10.0

20.0

30.0

40.0

50.0

.0 -.10 -.20 -.30 -.40

-100.0

-80.0

-60.0

-40.0

-20.0

.0 .0 -.10 -.20 -.30 -.40

-100.0

-80.0

-60.0

-40.0

-20.0

.0

EXPERIMENTAL FIBRE TRILINEAR

Curv.(m-1)

Mom.(kN.m) Mom.(kN.m)

Mom.(kN.m) Mom.(kN.m)

Curv.(m-1)

Curv.(m-1) Curv.(m-1)

a) LNEC beam section S2 b) KH beam section S2

M+

M –

M –

.0 .05 0

50

.025

Zoom

fsy

lp1( )

Page 206: Seismic Assessment of Reinforced Concrete Frame Structures

182 Chapter 4

higher than those expectable for no pull out contribution. For both beam types, the numeri-

cal result deviations from experimental values of are more apparent for the negative

direction than for the positive one (for in the LNEC beam case, a zoom view is pro-

vided up to the near-yielding zone, in order to confirm that deviation), which is related to

the different contents of tensioned steel bars inside the same width (25cm): for only

are engaged, while are resisting the applied moment , thus leading to

higher tensile force to be transferred to the concrete in poorer bond conditions (due to the

increased density of reinforcement) and, consequently, inducing larger slippage.

• The LNEC beam sections exhibit rather different post-yielding behaviour in the two bend-

ing directions due to the slab effect. For the positive direction, compressive stresses can

spread along the slab flange within a reduced thickness, meaning that the internal force

lever arm tends to increase for curvature values above . This effect, combined with the

strain hardening of steel, leads to the significant increase of resisting moment until crushing

and spalling of the unconfined concrete cover occurs; from then on, the internal resistance

starts to decrease and tends to the ultimate moment. By contrast, for , the increase of

steel force due to hardening is counterbalanced by the reduction of internal lever arm

caused by the greater neutral axis depth required to generate the compressive force neces-

sary to satisfy equilibrium; therefore, the hardening effect in almost vanishes and the

drop of resistance due to spalling becomes much less apparent, yet still perceptible. The

fibre analysis can approximately trace out these results although anticipating the effect of

concrete cover spalling, because, rather than assuming the typical post-yielding plateau, the

steel hardening is considered starting immediately after . Thus, in the fibre modelling,

forces developed in the yielded steel bars are higher than in the experiment, requiring larger

compressive forces in the concrete and leading to increased moments; this means that

crushing and spalling of concrete cover is triggered off (in the fibre analysis) for curvature

values lower than in the experiment, quite apparent for and also detectable for .

• On the contrary, the behaviour of KH beams is quite similar for both loading directions and,

besides the already referred deviations of yielding (and post-yielding) curvatures due to pull

out effects, the numerical analyses provide good approximations well within the scatter of

material properties. In comparison to LNEC beams, it is worth noting the significant differ-

ence of (because in KH beams there is no slab reinforcement contribution) and the

lower hardening effect for in KH beams (explained by reasons similar to those for the

LNEC beam behaviour in the negative direction).

ϕy

M+

M+

2φ12 4φ12 M -

ϕy+

M -

M -

εsy

M+ M -

M -

M+

Page 207: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 183

The above comments allow to accept that, within the limitations inherent in the trilinear curve,

it actually provides a good approximation of the experimental moment-curvature relationship;

however, it is apparent that, for a better fitting of the post-yielding/pre-spalling behaviour, at

least one further branch would be suitable to enhance the adequacy of a multi-linear step wise

approximation, particularly if the slab contribution to the strength is to be duly accounted for.

Results of numerical simulations using the flexibility element for monotonic tests V1 and V2

are shown in Figure 4.20 together with the experimental ones.

Figure 4.20 LNEC and KH beams S2: monotonic tests V1 and V2

EXPERIMENTAL TRILINEAR

Force (kN)

Displ.(m)

a) LNEC beam S2 b) KH beam S2

.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

.0 .02 .04 .06 .08 .10 .12

.0

-10.0

-20.0

-30.0

-40.0

-50.0

-60.0

-70.0

.0 -.02 -.04 -.06 -.08 -.10 -.12 .0

-10.0

-20.0

-30.0

-40.0

-50.0

-60.0

-70.0

.0 -.02 -.04 -.06 -.08 -.10 -.12

.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

.0 .02 .04 .06 .08 .10 .12

Force (kN)

Displ.(m)

Force (kN)

Displ.(m)

Force (kN)

Displ.(m)

V1

V2

V1

V2 .0

-70.0

.0 -.01 -.02

Zoom

Page 208: Seismic Assessment of Reinforced Concrete Frame Structures

184 Chapter 4

From the comparison of numerical and experimental force-displacement curves for both

LNEC and KH beams (Figure 4.20), the following can be observed:

• Results agree reasonably well, with exception of the test V1 on the LNEC beam due to the

slab effect already explained.

• Yielding forces are well estimated, thus fully agreeing with what has been said for the M-ϕ

curves.

• The underestimation of yielding displacements, caused by the missing contribution of the

pull out effect, is quite apparent. To some extent, this fact does not allow an adequate com-

parison of the numerical and experimental post-cracking branches, which would be desira-

ble in order to find out the effect of the cracking plateau and the cracking development in

the global response. Nevertheless, a zoom view of the post-cracking stage is included for

the test V2 of the LNEC beam, from which it can be observed that, despite a somewhat

underestimated yielding force, the post-cracking branch would tend to the experimental one

if the pull out contribution were included.

The results of cyclic test simulations are shown in Figures 4.21 and 4.22, respectively for

LNEC and KH beams. Comparison with experimental diagrams is included, not only in terms

of force-displacements, but also for moment - average curvature in the first plastic zone, which

required the average curvature from numerical analysis to be obtained by integration in the

plastic length in order to have consistent results. Due to problems in the instrumentation

setup, part of the experimental results for the LNEC beam - test V6 were not available; the cor-

responding part is indicated in Figures 4.21-b) by a dashed line.

Results suggest the following comments:

• The lack of pull out contribution is still apparent at the yielding stage; however, for further

cycling, the less good simulation of this effect becomes less relevant in presence of other

phenomena such as the pinching due to reinforcement asymmetry and the unloading/reload-

ing stiffness deterioration.

• The discrepancy of positive post-yielding strength for the LNEC beam, so apparent in the

monotonic test V1, actually vanishes for cyclic loading. This is related with the loading and

yielding occurrence for , prior to yielding for , which has induced significant crack-

ing in the slab and yielding in the top reinforcement. In the subsequent cycles for reverse

load direction , the slab cracks could not close again (as observed in the experiment

(Pipa (1993))), which, combined with the Baushinger effect in the top steel, led to signifi-

lp1

M - M +

M +( )

Page 209: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 185

cant curvature development upon yielding for and, consequently, the hardening effect

in the positive direction became less pronounced than for monotonic loading conditions.

Therefore, it can be concluded that, for cyclic conditions, even if the apparent hardening

(with significant contribution from the compressed slab participation) happens to develop

for the very first cycle with the slab under compression, for subsequent cycles, it vanishes

as a result of the described effects.

Figure 4.21 LNEC beam S2: tests V5 and V6

M +

a) Test V5: Force-displacement and Moment-Curvature diagrams

Force (kN)

Displ.(m)

Mom. (kN.m)

Curvat.(m-1) -.08 -.06 -.04 -.02 .0 .02 .04 .06

-70.0

-60.0

-50.0

-40.0

-30.0

-20.0

-10.0

.0

10.0

20.0

30.0

-.40 -.30 -.20 -.10 .00 .10 .20 -100.0

-80.0

-60.0

-40.0

-20.0

.0

20.0

40.0

b) Test V6: Force-displacement and Moment-Curvature diagrams

Instrumentation fault Instrumentation fault

Force (kN)

Displ.(m) -.08 -.06 -.04 -.02 .0 .02 .04 .06

-70.0

-60.0

-50.0

-40.0

-30.0

-20.0

-10.0

.0

10.0

20.0

30.0 Mom. (kN.m)

Curvat.(m-1) -.40 -.30 -.20 -.10 .00 .10 .20

-100.0

-80.0

-60.0

-40.0

-20.0

.0

20.0

40.0

V5

V6

EXPERIMENTAL TRILINEAR

Page 210: Seismic Assessment of Reinforced Concrete Frame Structures

186 Chapter 4

Figure 4.22 KH beam S2: tests V5 and V6

• As for the monotonic tests, the yielding strength is well estimated.

• For the level of involved deformations, the rule for unloading stiffness degradation seems to

be adequate.

• Reloading towards slab under compression is done with excessively reduced stiffness in the

numerical simulation (see Figures 4.21-a)). Indeed, the rule of aiming at the previous maxi-

mum point leads to strong deterioration of this reloading stiffness, at least for the first cycles

EXPERIMENTAL TRILINEAR

a) Test V5: Force-displacement and Moment-Curvature diagrams

b) Test V6: Force-displacement and Moment-Curvature diagrams

-50.0

-40.0

-30.0

-20.0

-10.0

.0

10.0

20.0

30.0

-80.0

-60.0

-40.0

-20.0

.0

20.0

40.0

-50.0

-40.0

-30.0

-20.0

-10.0

.0

10.0

20.0

30.0

-80.0

-60.0

-40.0

-20.0

.0

20.0

40.0

Force (kN)

Displ.(m)

Mom. (kN.m)

Curvat.(m-1) -.08 -.06 -.04 -.02 .0 .02 .04 .06 -.40 -.30 -.20 -.10 .00 .10 .20

Force (kN)

Displ.(m)

Mom. (kN.m)

Curvat.(m-1) -.08 -.06 -.04 -.02 .0 .02 .04 .06 -.40 -.30 -.20 -.10 .00 .10 .20

V5

V6

Page 211: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 187

at a given deformation level; for subsequent cycles with the same displacement, the stiff-

ness tends to approach the experimental one as exhibited in Figures 4.21-b).

• The stiffness for reloading with pinching effect is also overly reduced. Actually, the pinch-

ing moment (i.e., that for the crack closure point) is well approximated but the correspond-

ing deformation appears over-estimated (see, for instance, Figures 4.21-a)), thus

excessively reducing the first reloading branch stiffness. Additionally, for larger displace-

ments (say 3% to 4% of the member length), the pinching effect can be just roughly cap-

tured by two straight lines, and, rather than closely following the force-deformation curve,

the approximation should preferably aim at compensation of loop areas (i.e., dissipated

energy).

• Strength degradation in the negative direction is strongly influenced by the pinching effect.

In fact, due to the excessively low reloading stiffness induced by pinching, the obtained

resistance for imposed deformation becomes lower than the experimental one (see Figures

4.21-b) and 4.22-b), thus, somehow overlapping the appearance of strength degradation.

Instead, for the positive direction, low degradation occurs and, where visible (ex. Figures

4.21-b)), the numerical results lead to less degradation than the experiment; this may be

related with, either the lower dissipated energy or a low value of the strength degradation

parameter, or even with both aspects.

Attention is drawn to the similar shape of force-displacement and moment-curvature diagrams;

indeed, such similarity was expected since the non-linear response is mainly controlled by that

plastic zone behaviour, but it means also that corrections on the reloading stiffness and pinch-

ing modelling at the section level will have straight correspondence at the global level. In other

words, should improvements be undertaken to have results closer to the experimental ones, it

appears to be sufficient that they are done in the section model.

4.3.4 Remarks on model validation results

The above presented results show that, overall, quite reasonable numerical simulations can be

obtained with the flexibility element formulation under monotonic and cyclic conditions.

However, deviations do occur, particularly concerning the cyclic loading cases, which are

related to limitations or less adequate rules of hysteretic behaviour at the section level, with

particular emphasis for the modelling of reloading stiffness, either with or without pinching

effect included.

Page 212: Seismic Assessment of Reinforced Concrete Frame Structures

188 Chapter 4

As already mentioned, the flexibility element formulation was developed herein having the

main requirement that the model be based on a trilinear skeleton curve and on multi-linear step

wise features for hysteretic behaviour simulation. However, no specific restrictions were put

for hysteretic rules, which means that improvements in the section model are perfectly compat-

ible with the flexibility formulation and may be sufficient to achieve results closer to the exper-

imental ones. Furthermore, it is reminded that the major concern of the present work is more

the global element formulation than the local section model refinement, yet keeping the formu-

lation open to new improvements at the section level. For these reasons and, despite the above

highlighted limitations, the presently adopted Takeda-type model was still kept for further cal-

culations. In this context, the present validation examples serve mainly for global assessment

of the model and for identifying the major limitations to bear in mind when analysing other

cases of complete structures in the following chapters.

Finally, it is noteworthy that the validation of a global element formulation is much harder,

cumbersome and even less complete than a section model validation. Actually, it is very diffi-

cult to predict and to generate the whole range of situations likely to develop in a given ele-

ment, because it is strongly dependent on several factors, viz the moment distribution, the

cracking and yielding development, the state of end sections which directly affect the corre-

sponding plastic zone state and rarely behave independently. Even more difficult is the availa-

bility of single elements or sub-assemblages experimentally tested that had gone through

several possible cases of the element state; typically, cantilever or simply supported members

are tested under simple loading conditions not likely to generate complex internal force distri-

butions and load reversals as often developed in members integrated in complete frame struc-

tures.

Therefore, part of the validation process had to be done by artificially generating various load-

ing conditions in single elements and by checking the output of numerical analyses for reason-

able results (examples: bilinear moment distributions, possibly reaching full-cracking of one or

both element parts; loading, unloading and reloading sequences affecting the plastic zone

development and aiming at generating cases as those included in Figures 3.23 and 3.24).

However, none of such tedious local validation analyses is included herein; instead, attention is

drawn to the validation at the global structure level as included in Chapter 5, which is deemed

more meaningful in the context of global element modelling.

Page 213: Seismic Assessment of Reinforced Concrete Frame Structures

NUMERICAL IMPLEMENTATION, AUXILIARY TOOLS AND VALIDATION 189

4.4 Conclusions

In this chapter the most significant tasks for implementation of the flexibility element were

addressed and a new algorithm was presented for the definition of trilinear approximations of

moment-curvature relationships. Additionally, several isolated members were analysed with

the developed formulation and implemented tools, as part of the validation process at the sin-

gle element level, and results were discussed in comparison with experimental ones.

Implementations were carried out in the computer code CASTEM2000, for which a prelimi-

nary review of the basic features was included and complemented with a simple but illustrative

example of structural analysis.

CASTEM2000 operators affected by the implementation of the flexibility formulation were

presented and briefly described; the major adaptations to accommodate the new element for-

mulation were generally reported.

For the definition of moment-curvature trilinear approximations, an auxiliary tool has been

implemented as a new CASTEM2000 operator, and based on an efficient algorithm for the

analysis of rectangular and T-shape reinforced concrete sections. It has been designed to over-

come the need of fibre discretization, yet accounting for realistic material behaviour models:

steel is assumed behaving bilinearly (with strain hardening) and concrete follows a parabolic/

linear-softening/residual-plateau model type, either in confined or unconfined conditions.

While cracking points are defined by closed-form expressions, the non-linear range turning

points (viz yielding and ultimate) are treated by a common process and the distinct features

reduce to the point definition criterion; such a process offers the additional possibility of con-

sidering supplementary points by simply providing adequate definition criteria.

In a restricted context of the validation of the flexibility element model using experimentally

tested specimens, rectangular and T-shape section cantilever beams were analysed by one flex-

ibility element, both in monotonic and in cyclic loading conditions. Results were compared

with experimental ones, showing that good numerical simulations can be obtained with the

present formulation. However, modelling limitations were found, which are related to the hys-

teretic rules at the section level, namely the reloading stiffness modelling, with or without the

pinching effect. It has been recognized that corrections for such limitations are essentially

localized in the section model and do not interfere with the element formulation. However,

Page 214: Seismic Assessment of Reinforced Concrete Frame Structures

190 Chapter 4

such corrections were not performed herein and, therefore, the above referred limitations have

to be taken into account in the numerical analyses of the next chapters.

Page 215: Seismic Assessment of Reinforced Concrete Frame Structures

Chapter 5

THE 4-STOREY FULL-SCALE BUILDING

TESTED AT ELSA

5.1 Introduction

The present chapter deals with the experimental research related to the four-storey full-scale

reinforced concrete building tested in the European Laboratory for Structural Assessment

(ELSA) of the Joint Research Centre (JRC) at Ispra, Italy.

The initiative was included in the framework of a Cooperative Research Programme on the

Seismic Response of Reinforced Concrete Structures (Carvalho (1991,1992,1993)) as part of

the activities of the European Association of Structural Mechanics Laboratories (EASML).

The work carried out by the so-called “Reinforced Concrete Working Group” of the EASML

started with the first phase (Carvalho (1991)) in which: a) similarities and differences were

identified between seismic design and analysis methods in different european countries and, b)

numerical studies were performed concerning the design of some buildings using the current

drafts of Eurocode 2 (1991) (EC2) and Eurocode 8 (1994) (EC8) and concerning also non-lin-

ear analysis with artificially generated accelerograms consistent with the EC8 elastic spectrum.

The objective of the second phase (Carvalho (1992,1993)) was further assessment of the EC8

and the definition of damage indicators and failure criteria for plastic zones of structural mem-

bers. Simultaneously, more emphasis was put on testing activity, both at the member and the

complete structure levels. In this line, a full-scale high-ductility structure was designed accord-

ing to EC2 and EC8, to be pseudo-dynamically tested in the ELSA laboratory in order to allow

a detailed comparison between the actual and the intended design behaviour.

Page 216: Seismic Assessment of Reinforced Concrete Frame Structures

192 Chapter 5

The specific features of the Pseudo-Dynamic (PSD) method, as a hybrid numerical-experimen-

tal testing procedure, are briefly described in 5.2, including references to time integration tech-

niques (a key issue in the PSD method) and to substructuring procedures in PSD tests.

Examples of PSD testing activity in the ELSA laboratory are also provided.

The main topics concerning the structure design and layout are given in 5.3, while the charac-

terization tests of concrete and steel properties are referred in 5.4, along with the reduced scale

member tests performed in complement of the full-scale ones.

The full-scale testing activity in the four-storey building is described in 5.5. After brief refer-

ences to the specific PSD test setup, the used input accelerogram and the preliminary tests for

dynamic characterization of the structure, the main emphasis is put on the results of the seismic

tests and the final cyclic tests. Seismic tests were performed for two levels of earthquake inten-

sity on the bare structure and additional tests were carried out for two mansonry infilled frame

configurations. Final cyclic tests were performed on the bare frame structure aiming at a near

failure stage.

Finally, the main conclusions of the whole experimental activity related to the four-storey

building are summarized in 5.6.

5.2 The Pseudo-Dynamic test method. An overview

5.2.1 General

The Pseudo-Dynamic method is a hybrid testing method which combines classical experimen-

tal techniques with on-line computer simulation of structural dynamic behaviour (Donea et al.

(1990)). Displacements are quasi-statically imposed to the structure, satisfying the discrete

equations of dynamic equilibrium at each time step, and the structure restoring forces are

experimentally measured.

Since the inertia forces are analytically modelled, in a PSD experiment the expected real veloc-

ities do not need to be reproduced as it is required in shaking table tests. Therefore, the demand

for hydraulic flow power is much lower, which allows for testing structures at a larger scale

than in shaking tables. This is quite adequate for the analysis of structures made out of materi-

als, such as reinforced concrete, whose behaviour is influenced by aggregate size, micro-crack-

Page 217: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 193

ing, bond between concrete and steel, etc., which are phenomena difficult to simulate in small

scale models. Moreover, the time scale being substantially amplified (about two to three orders

of magnitude), it allows for a more detailed inspection of damage evolution during the test.

Also, the ability to use substructuring techniques in PSD testing (Buchet and Pegon (1994))

makes it very attractive for structures in which, some parts can be easily modelled numerically,

while other parts are simulated by a physical specimen, only where strictly necessary. This is

the case of bridges, where the non-linear structural response to horizontal seismic forces is

mainly controlled by the piers, while the deck (difficult to fit inside the laboratory) can be ade-

quately modelled by numerical tools. Thus, the piers are physically tested, the deck is numeri-

cally simulated and adequate interface conditions are taken into account.

On the other hand, the reduced velocity of load application may have some drawbacks. If the

structure constitutive materials are strain-rate sensitive, the restoring forces for a given dis-

placement level may actually differ from those in a real time scale experiment, such as shaking

table tests. Fortunately, this seems not to be the case of reinforced concrete for which the rate

sensitivity appears of reduced importance as reported by Gutierrez et al. (1993).

Additionally, the PSD testing of structures with distributed mass is not a straightforward task,

since the inertia forces also develop in a distributed manner and would demand a similar distri-

bution of actuators. Such a requirement cannot be practically met and, therefore, more elabo-

rated PSD testing techniques and further assumptions are needed for distributed mass systems;

by contrast, shaking table testing does not present this limitation. However, for building frame

structures this problem is meaningless because the mass is typically concentrated on the floors.

It turns out that these two testing techniques appear as complementary and the adequacy of

each of them to simulate the dynamic structural response has to be judged in accordance with

the specific type of structure and material under analysis.

5.2.2 Time integration techniques

In the PSD method the non-linear behaviour modelling is pursued by means of the measure-

ment of restoring forces due to imposed displacements. The external loading, usually ground

accelerations, is incrementally applied and the dynamic equilibrium equations are solved by an

adequate integration algorithm, the inertia and damping forces being analytically modelled.

Page 218: Seismic Assessment of Reinforced Concrete Frame Structures

194 Chapter 5

The use of step-by-step integration methods is recalled in Appendix E within the context of

numerical non-linear analysis for the integration of

(5.1)

which is achievable by considering additional expressions specific to the adopted method (see

Eqs. (E.5) and (E.6)). In the PSD algorithm, Eq. (5.1) is used together with such expressions

(to be discussed below), in order to perform the following tasks:

• computation of displacements according to the integration method expressions;

• application of to the structure and measurement of restoring forces in the actuators, lead-

ing to ;

• calculation of and using Eq. (5.1), along with other expressions specific to the inte-

gration method, and restart for the next step.

Time integration methods proposed for PSD testing consist of explicit, implicit and hybrid

algorithms. As referred in Appendix E, for the first method, any displacement prediction for a

given step is based only on the previous step response, while for the implicit algorithms it also

depends on the response of the step under analysis.

The explicit methods, although leading to good results, have the major drawback of condi-

tional stability. This means that the adopted time step must be smaller than a threshold related

with the highest frequency of vibration, in order to achieve stability. For certain structures this

implies a greater number of steps to adequately lower the time interval, which may increase the

errors due to testing control.

By contrast, implicit methods are unconditionally stable, which makes them rather attractive

for testing very stiff structures, or with many degrees of freedom (d.o.f.). However, for struc-

tures with non-linear behaviour, either the tangent stiffness matrix must be obtained, or an iter-

ative procedure for displacement correction has to be adopted. For this reason, implicit

algorithms have been considered not usable in the PSD context, since, on the one hand the

“tangent stiffness matrix” is very difficult to assess experimentally, and, on the other hand, due

to the dependence of structural deformations on the loading history, any corrective iterative

procedure would introduce spurious hysteretic cycles.

However, during the last years, procedures have been proposed to adopt implicit algorithms in

M ak⋅ C vk⋅ rk+ + qk=

dk

dk

rk

ak vk

Page 219: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 195

PSD testing, mainly based, either on slight modifications at the experimental level, or in the

application of different numerical expressions for the displacements.

In the following sections, details are given on the use and implementation of classic explicit

and implicit algorithms, and references are made to other recent proposals consisting on hybrid

explicit-implicit techniques.

5.2.2.1 Newmark explicit algorithm

The Newmark family of time integration methods is described in Appendix E, according to

which the integration of Eq. (5.1) is based on Eqs. (E.5) and (E.6), re-written here for conven-

ience

(5.2)

(5.3)

The central difference method is obtained by choosing and ; in this case the

stability condition for such explicit algorithm is which must be satisfied for the

higher natural frequency associated with the adopted structural discretization.

The displacements are therefore given by

(5.4)

and the implementation of the central difference algorithm in the PSD method involves meas-

uring the restoring forces , followed by the computation of the acceleration and velocity

vectors by Eqs. (5.2) and (5.1), leading to

(5.5)

(5.6)

At this stage the response for is known and the algorithm proceeds to the next step.

vk vk 1– 1 γ–( )ak 1– γak+[ ]∆t+=

dk dk 1– vk 1– ∆t 12--- β–⎝ ⎠⎛ ⎞ ak 1– βak+ ∆t2+ +=

β 0= γ 1 2⁄=

0 ωn∆t 2≤ ≤

ωn

dk dk 1– ∆tvk 1–∆t2

2-------ak 1–+ +=

rk ak

vk

M ∆t2-----C+⎝ ⎠

⎛ ⎞ ak qk rk– C vk 1–∆t2-----ak 1–+⎝ ⎠

⎛ ⎞–=

vk vk 1–∆t2----- ak 1– ak+[ ]+=

tk

Page 220: Seismic Assessment of Reinforced Concrete Frame Structures

196 Chapter 5

In Figure 5.1 a schematic illustration helps to clarify the procedure for practical implementa-

tion of the explicit PSD method.

Figure 5.1 Implementation of the explicit PSD method.

Displacements are continuously applied following a pre-defined ramp from until .

The effectively applied displacements are obtained as a feed-back signal from the displace-

ment transducers and, if they do not match the desired displacements , the correction

is enforced as a new input signal for the actuators. When is satisfied, the

step is complete and information is sent to the main computer in order to process the next

step.

5.2.2.2 α-implicit algorithm

The impossibility of simultaneously achieving second order accuracy and numerical dissipa-

tion capabilities with the unconditionally stable implicit Newmark method is pointed out in

Appendix E.

In order to overcome this limitation, Hilber et al. (1977) proposed another implicit method by

introducing a new parameter in Eq. (5.1) which allows for numerical dissipation. Eqs. (5.2)

and (5.3) remain unchanged, whereas Eq. (5.1) becomes

(5.7)

which reduces to Eq. (5.1) when . According to the authors, if , and are taken as

Controller

dk

dk 1– tdk

Ramp Generator

ed

+ -x

dk

rk

Main Computer

Calculate:dk

ak

Control

Algorithm

Servo-Valve

Actuator

Structure

+

Compute:

Set:dk 1– dk=vk 1– vk=ak 1– ak=

vk

d dk 1– dk

x

d

e d x–= x dk=

tk

α

Mak 1 α+( )Cvk αCvk– 1 α+( )rk αrk 1––+ + 1 α+( )qk αqk 1––=

α 0= α β γ

Page 221: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 197

(5.8)

an implicit, unconditionally stable and second order accurate algorithm with numerical dissi-

pation can be obtained.

In the PSD method context, the use of this integration algorithm has the major problem of dis-

placement dependency on the unknown accelerations . However, Thewalt and Mahin

(1991) proposed a solution for this problem by means of a slight modification on the testing

sequence, such that information already existing in the analogue data flow is used to apply the

implicit displacements .

In order to explain this technique, let Eq. (5.2) be substituted in Eq. (5.7) so that accelerations

can be written as follows

(5.9)

where

(5.10)

consist of terms independent of the response for . The velocity vector can be given by

(5.11)

and replacing Eq. (5.9) in Eq. (5.3) leads to

(5.12)

where

(5.13)

13---– α 0<≤ γ 1 2α–( )

2---------------------= β 1 α–( )2

4-------------------=

dk ak

dk

ak M11– qk α+ αrk 1– Cvk– 1 α+( )rk–+[ ]=

qk α+ 1 α+( )qk αqk 1––=

vk vk 1– 1 γ–( ) 1 α+( )∆tak 1–+=

M1 M γ 1 α+( )∆tC+=

tk

vk vk γ∆tak 1–+=

dk dk Brk–=

dk dk 1– ∆tvk 1–∆t2

2------- 1 2β–( )ak 1– ∆t2βM1

1– qk α+ αrk 1– Cvk–+( )+ + +=

B ∆t2βM11– 1 α+( )=

Page 222: Seismic Assessment of Reinforced Concrete Frame Structures

198 Chapter 5

Note that Eq. (5.12) is implicit on displacements since the restoring forces appearing on

the second term depend on . In a numerical analysis procedure, can be expressed in terms

of by means of the tangent stiffness matrix. In the PSD test context, Thewalt and Mahin

proposed the following procedure: the explicit portion of displacements is digitally com-

puted and the corresponding signal is sent to the controllers to be applied by actuators; the

remainder is calculated in analogue form using the feedback voltages from specimen

restoring forces which are continuously measured in the load cells.

The practical implementation of this technique is illustrated in Figure 5.2 and can be described

as follows: the target displacements are defined as and computed by the first expression of

Eqs. (5.13). While in the conventional explicit method is progressively applied in ramp

form until the actually achieved displacements reach , in this implicit method the driving

signal for the controller is now modified at the analogue level to be ,

where is the restoring force which is continuously measured and is constant as given by

the second expression of Eqs. (5.13). At the end of the step, i.e., when the corrective signal

vanishes, the measured displacements match corresponding to the implicit

ones.

Figure 5.2 Implementation of the implicit PSD method.

Attention must be drawn to the fact that displacements achieved at step completion depend

on the measured restoring forces and are unknown to the main computer beforehand. During

displacement application the information updating is kept at the analogue level and is not seen

dk rk

dk rk

dk

dk

Brk–( )

dk

dk

x dk

d x– e d x Br+( )–=

r B

e

x dk Brk–= dk

Controller

dk

dk 1– tdk

Ramp Generator

ed+

- x

dk

rk

Control

Algorithm

Structure

d

Br

+-

Servo-Valve

Actuator+

Main Computer

Calculate:dk

akCompute:

Set:dk 1– dk=vk 1– vk=ak 1– ak=

vk

dk

Page 223: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 199

by the computer which makes error checking more difficult than in the explicit PSD method.

It is noted that this has been the first implicit technique successfully applied in the PSD testing

of structures with several d.o.f.. However it has been achieved through modifications at the

system hardware level, which tends to limit its use in laboratories traditionally only equipped

for the explicit PSD method.

The error checking and implementation difficulties of this implicit PSD method has steered the

research during the last decade towards mixed explicit-implicit algorithms. In the next para-

graphs, the main topics and advantages of these algorithms are highlighted.

5.2.2.3 Mixed explicit-implicit algorithms

The basic idea of mixed methods is that the simplicity of the explicit PSD version is retained

and combined with the unconditional stability of an implicit method.

A typical example of these methods is the Operator-Splitting (OS) method originally proposed

by Nakashima et al. (1990) and based in an algorithm developed by Hughes et al. (1979), in

which the restoring force vector is approximated by the sum of one elastic term with a nonlin-

ear one. For convenience Eq. (5.1) is re-written in the form

(5.14)

where is the elastic stiffness matrix, is the updated stiffness matrix (tangent or secant,

depending on whether Eq. (5.14) refers to incremental or to total quantities) and is a dis-

placement predictor based only on the previous step response. Denoting by the

restoring forces due only to the predictor , and comparing Eqs. (5.14) and (5.1), it can be

concluded that are estimated by and corrected by the difference between target and pre-

dicted displacements through the elastic stiffness.

Eq. (5.14) can be also re-arranged in the form

(5.15)

which suggests its application to the PSD method as follows:

• the elastic matrix is first obtained, either numerically or experimentally;

Mak Cvk Kdk Ke dk dk–( )+ + + qk=

Ke K

dk

rk Kdk=

dk

rk rk

Mak Cvk Kedk rk Kedk–( )+ + + qk=

Page 224: Seismic Assessment of Reinforced Concrete Frame Structures

200 Chapter 5

• the displacement predictor is applied to the structure and the corresponding restoring

forces are measured;

• the term inside brackets can be evaluated and rearranged in the second member, leading to

an implicit equation that can be solved numerically (because is known and constant) to

obtain the corrected displacements .

This method has been presented in the context of the Newmark family integration algorithms

and, in particular, for and the classical Newmark method (with trapezoi-

dal rule for acceleration) is recovered. According to the authors, the method is unconditionally

stable, as long as the structural non-linearity is of the softening type, and with an accuracy

equivalent to that of the classical Newmark method at lower response modes.

If numerical dissipation is to be included, this method can be also used together with the above

referred -implicit algorithm as proposed by Combescure and Pegon (1994) leading to the so-

called -OS.

Following the option of using the PSD only in the explicit form, Shing et al. (1990) proposed

another method in which a numerical iterative scheme is used to apply the implicit part of the

displacements, instead of the analogue correction due to the restoring forces as proposed by

Thewalt and Mahin. Such iterative procedure is based on the initial stiffness matrix and, in

order to prevent spurious hysteretic cycles, a reduction factor is used for displacements which

reduces the risk of “overshooting”, i.e., applying displacements larger than the desired ones.

Further details about this method are out of the scope of this work, but can be extensively

found in Shing et al. (1990), where the method accuracy is proved with an error-propagation

analysis and numerical dissipation is shown to be activated due to an error-correction method

provided in the process.

5.2.3 Substructuring in the PSD method

Due to the mixed numerical-experimental character of the PSD method it has been possible to

implement substructuring techniques, through which part of the structure is analytically mod-

elled, whereas the other part is physically simulated and tested.

Such improvement has considerably opened the field of the PSD method applications, namely

to the analysis of very large structures (for instance, bridges (Pinto et al. (1996))), to the possi-

dk

rk

Ke

dk

β 1 4⁄= γ 1 2⁄=

α

α

Page 225: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 201

ble simulation of asynchronous motions along structure supports (Pegon (1996)) as well as

soil-structure interaction effects, etc. Although the substructuring technique is beyond the pur-

pose of this work, a brief reference is made to the main features involved.

The use of substructuring in PSD testing requires an adequate behaviour model for the numer-

ically simulated part and an efficient time integration technique. It is noted that the integration

method assumes an increased importance in the PSD testing with substructuring, since this

usually corresponds to large numbers of tested and modelled d.o.f.. To this aim the -OS

method mentioned in 5.2.2 combines the essential requirements in this type of PSD testing,

namely: i) the unconditional stability, ii) a selective numerical damping only for the highest

frequency modes and iii) the non-iterative feature for which the implementation simplicity of

the explicit central difference method is still valid.

During a PSD test with substructuring two processes are running in parallel, corresponding to

the numerical and to the tested substructures, between which a data flow is established at pre-

cise moments of each load step. For clarity sake the whole structure d.o.f. are divided into three

types: those of the numerical substructure (SS), those of the tested substructure (TT) and those

of the interface (ST) between both substructures.

The whole set of d.o.f. is statically condensed to ST+TT and the PSD test is run only over these

d.o.f. following the usual algorithm, provided that the influence of the above condensation is

adequately included in the matrices of the process (see Buchet and Pegon (1994)). In the pre-

dictor phase, the displacement vector is split into sub-vectors corresponding to the SS and to

the ST+TT d.o.f. and sent, respectively, to the numerical substructure and to the tested one. In

each of them the associated restoring forces are evaluated and sent back, together, to provide

information for the solution of the condensed equilibrium equation in the ST+TT d.o.f.. Once

this task is concluded the whole response can be updated for the step under analysis, both in

the numerical and the tested substructures, and the process is ready to proceed to the next step.

It is noted that, with such an implementation scheme, only slight modifications are necessary

in the explicit PSD method, which are mainly related with information interchange between

both substructures. This contributes to the implementation simplicity in laboratories tradition-

ally equipped for the PSD test without substructuring.

α

Page 226: Seismic Assessment of Reinforced Concrete Frame Structures

202 Chapter 5

5.2.4 Applications of the PSD method at the ELSA laboratory

Since 1992 the ELSA laboratory is performing PSD tests for the seismic assessment of struc-

tures, already covering a wide range of different structural systems.

The testing activity started with a full scale moment resisting steel frame designed with EC8

(Kakaliagos (1994)) which helped to fix the test setup and proved the laboratory capabilities

for PSD testing pointed out by Donea et al. (1996). Besides the global performance of the

structure a particular aspect under investigation was the behaviour of semi-rigid beam-column

joints; details about this tests can be found in Kakaliagos (1994).

The subsequent full-scale tests were performed in the four-storey reinforced concrete structure

the present chapter refers to. These tests are extensively described and discussed in the next

sections, for which no further reference is made herein.

A broad PSD-testing campaign was undertaken in large scale RC bridge specimens, in the

framework of an integrated European programme of pre-normative research in support of EC8,

aiming at providing background and improvement for analysis and design methods. The

bridges under study had three piers with height ranging from 7 to 21 m and identical rectangu-

lar hollow cross-sections and a continuous deck 200 m long. The specimens at 1:2.5 scale were

tested with the substructuring technique, using physical models for the piers and numerical

modelling for the deck.

Two main configurations of regular and irregular bridges were considered and tested initially

for fixed support conditions and synchronous motion, with an artificial accelerogram and aim-

ing at the assessment of the influence of irregularity on the seismic ductile behaviour of bridge

piers. Later, other tests were carried out on specimens with isolation/dissipation devices in

order to check the possibility of obtaining a more homogeneous ductility demand for the piers

and in the last testing phase an asynchronous input motion was simulated at the base of the

piers; detailed information about these tests can be found in Pinto et al. (1996).

The seismic behaviour of monumental structures is increasingly drawing the attention of

researchers due to the need of preserving the historical heritage. Along this trend-line, PSD

testing of representative models of monumental structures has already started at ELSA, of

which the first specimen was a large scale (1:3) model of the facade of the Palazzo Geracci in

Page 227: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 203

Sicily, reproduced in the laboratory with masonry stone blocks similar to those of the original

structure. The tests were quite successful and allowed to check the ability of the PSD setup to

test structures with high initial stiffness; reports on this tests can be found in Anthoine (1997).

Another monumental stone masonry structure to be tested is a full-scale model of the São Vice-

nte de Fora Monastery in Lisbon, of which the cloisters were considered a representative part.

A set of four arches was built in the laboratory, also with masonry adequately chosen with sim-

ilar characteristics to that of the monastery, and special devices were required for the loading

system. A preliminary study involving numerical analysis and definition of the experimental

model is detailed in Pegon and Pinto (1996).

Other examples of specimens to be pseudo-dynamically tested are a four-storey composite

steel-concrete structure designed according to Eurocode 3 (1992) (EC3), Eurocode 4 (1992)

(EC4) and EC8, as well as a two-storey RC building to be infilled with panels of hollow brick

masonry in several configurations.

Besides the specific scope associated with each test, obvious additional advantages are

obtained concerning validation of numerical models. Numerical simulations are usually per-

formed to anticipate the structural behaviour of the specimens and quite often either new mod-

els are developed or existing ones are improved. Thus the output of these large-scale tests

provides an excellent means of calibration and assessment of numerical models.

5.3 Structure design and layout

The specimen under analysis consists of a full-scale RC frame structure, four storey high and

designed in accordance with EC2 and EC8. It is a high-ductility structure (according to EC8

classification) whose general layout is shown in Figure 5.3.

The structure is symmetric in the testing direction with two identical lateral frames and a

stronger central one. In the orthogonal direction it is asymmetric due to different span lengths,

leading to a slight irregularity, which was introduced to have a more realistic building and for

possible tests in this direction (Carvalho (1993)). External columns have 40cm x 40cm cross-

section, while the central one has 45cm x 45cm. All beams are 30cm wide and 45cm high and

the solid slab thickness is 15cm.

Page 228: Seismic Assessment of Reinforced Concrete Frame Structures

204 Chapter 5

Figure 5.3 General layout of the 4-storey RC building tested at ELSA (dimensions in metres)

The structure was cast in place with normal-weight concrete C25/30 as specified in EC2 and

B500S Tempcore steel rebars and welded meshes. In spite of violating the requirements of

EC8 for ductility class “High” structures, concerning the strain at failure and the ratio of tensile

failure strength to yield strength, this kind of steel was used since it is gaining market in

Europe (Carvalho (1993)). Strong arguments sustained its adoption in order to profit the oppor-

tunity of a large scale test to assess the adequacy of this steel for seismic resistant structures.

However, to some extent, such option may have prevented the main scope of the test to be

achieved, i.e., an experimental contribution for the assessment of EC8 design rules. Actually, it

seems more convenient to first judge about this material adequacy by means of sub-assem-

0.80

3.00

3.50

13.3

0

3.00

3.00

0.45

0.40 x 0.40

0.45 x 0.45

10.0

0

5.00

5.00

6.00 4.0010.00

0.15

0.40 x 0.40

0.45 x 0.45

0.40 x 0.40

Rea

ctio

n W

all

Testing Floor

TestingDirection

d(+)

Z

XY

Y

X Z

a) Elevation

b) Plan view

Page 229: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 205

blages tests, since the majority of such tests which are available do not include this steel and

little experimental evidence exists about its behaviour (specially concerning eventual rein-

forcement slippage in the concrete core).

Design was performed for typical loads (additional dead load of 2.0 kN/m2 for finishing and

partitions and live load of 2.0 kN/m2) and for high seismicity, assuming a peak ground acceler-

ation of 0.30g, a soil type B and an importance factor of 1 (importance category III). Since the

regularity requirements are fulfilled, both in plan and elevation, the design was performed for

high regularity and no reduction in the behaviour factor is due. For the ductility class High this

yields a behaviour factor of 5.

Two independent planar models were used for the design analysis and the simplified method

prescribed in EC8 was used to account for torsion effects. More details about design, as well as

reinforcement layout drawings can be found in Negro et al. (1994).

5.4 Material properties and reduced scale member tests

The characterization of concrete has been done by means of compressive tests on 150 mm side

cubes (Negro et al. (1994)), leading to the mean values listed in Table 5.1 for each casting

operation. Estimates of the mean cylinder compressive and tensile strengths are also included,

having been obtained from the cube strength as mentioned later in section 6.2.2. Note that the

values are clearly higher than 33 MPa, representative of a C25/30 concrete mean compressive

strength, and show a significant scatter between casting phases.

Table 5.1 Mean cube (fcm,cub) and cylinder (fcm) compressive and tensile (fctm) strengths of concrete

Casting operation of: fcm,cub (MPa) fcm (MPa) fctm (MPa)1st storey Columns 49.8 44.8 3.67

1st storey Beams 56.4 51.4 4.09

2nd storey Columns 47.6 42.6 3.50

2nd storey Beams 53.2 48.2 3.87

3rd storey Columns 32.0 27.0 2.37

3rd storey Beams 47.2 42.2 3.48

4th storey Columns 46.3 41.3 3.43

4th storey Beams 42.1 37.1 3.14

Page 230: Seismic Assessment of Reinforced Concrete Frame Structures

206 Chapter 5

For the steel, tensile strength tests were carried out for several diameters, representative of

those used in the specimen construction, the available results being given in Table 5.2. In par-

ticular, concerning strains, the unique available information refers to the ultimate strain meas-

ured in a reference length of (where stands for the bar diameter).

In order to support the preparation of the full scale tests, a set of reduced scale (1:2/3) monot-

onic and cyclic tests was performed on 24 RC cantilever beams of 1.50m span. These tests

were performed at LNEC (Pipa (1993)) and KH (Carvalho (1993)) and have been already

referred and used in 4.3. The specimen cross-sections were taken representative of typical situ-

ations in the whole four-storey structure, namely one type of sections for the external frame

beams and another type for the internal frame (Carvalho (1993)).

Two main issues were envisaged:

• Evaluation of the slab influence in the response of beams, for which the whole specimen set

was divided into one part with T-shape beams (tested at LNEC) in order to include a contri-

bution from the slab, and another part having identical cross-sections but with no flange

included (tested at KH).

• Investigation of the B500S Tempcore steel adequacy in the cyclic behaviour of reinforced

concrete elements, in view of its lower ductile and hardening capacities comparatively to

traditional hot-rolled steel (reductions of about 20% and 10% on ductility and hardening,

respectively, are reported by Pipa (1993)).

According to Carvalho (1993) and Pipa (1993) the main results of these tests can be summa-

rized as follows:

Table 5.2 Mean tensile properties of steel

Diameter(mm)

Area(mm2)

Yielding Stress(MPa)

Ultimate Stress(MPa)

Ultimate Strain(%)

6 29.2 566.0 633.5 23.5

8 51.4 572.5 636.1 22.3

10 80.3 545.5 618.8 27.5

12 113.1 589.7 689.4 23.0

14 153.3 583.2 667.4 22.7

16 199.2 595.7 681.0 20.6

20 310.0 553.5 660.0 23.1

26 517.2 555.6 657.3 21.6

5φ φ

5φ[ ]

Page 231: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 207

• Besides the expected increase in strength at the cracking stage, the influence of the slab

width in the T-shape beams was also found at failure due to the increased section asymme-

try, which was quite apparent in the more “pinched” force-displacement diagrams of LNEC

test results comparatively to those from KH. Actually, failure was induced by the buckling

of bottom reinforcement bars after having been subjected to large strain variations, to which

the slab contribution was twofold: i) an increase of tensile strains at the bottom bars, due to

a larger but less deep area of compressed concrete and, ii) a higher strength when the top

bars are tensioned, due to the presence of slab reinforcement, which also increases the com-

pressive forces in the bottom bars and, thus, their tendency to buckle.

• The lower ductile and hardening capacities of the B500S Tempcore steel did not appear to

affect the available ductility of the specimens. Indeed, on the one hand the failure was trig-

gered mostly by buckling phenomena rather than the full exploitation of the steel ductile

capacity, and, on the other hand, the obtained plastic hinge length (mainly controlled by the

steel hardening) seemed to compare well with results of other experiments made with steel

having higher hardening capacity. However, the higher strength of the used steel indirectly

affects the greater tendency of bars to buckle, since, for the same tensile force, bars with

smaller diameter can be used.

• Quite evident pull-out effects were found due to bar slippage from the footing, contributing

to an increase of about 40% to 50% for displacements measured at yielding, relatively to

those expected if no pull-out had occurred. It must be pointed out that, the stronger the steel,

the greater the tendency of pull-out phenomena to occur, due to a lower lateral surface area

of the bars for the same tensile force to be transmitted to the concrete. Therefore, RC mem-

bers made with B500S Tempcore steel are, expectably, more prone to pull-out than those

made with lower yielding strength steel.

The overall conclusion from these tests highlighted the importance of taking the slab contribu-

tion into account in the beam response and sustained the compatibility of B500S Tempcore

steel with the EC8 requirements for ductile design of earthquake resistant structures; however,

the lack of more experimental evidence on the latter topic still recommends further research on

issues such as the anchorage conditions and the buckling phenomena.

5.5 Full-scale tests

The experimental programme related with the four-storey RC structure consisted on the fol-

Page 232: Seismic Assessment of Reinforced Concrete Frame Structures

208 Chapter 5

lowing phases (Negro et al. (1994), Negro et al. (1995)):

• Preliminary tests aiming at the dynamic characterization of the structure and the calibration

of the pseudo-dynamic test setup before the main non-linear tests.

• Seismic tests in the bare structure performed first for a low intensity level and then for a

high level of intensity (respectively, for 0.4 and 1.5 times the reference earthquake).

• Seismic tests on the structure with the external frames infilled with unreinforced masonry,

first in the totally infilled configuration and then without any infills on the first storey in

order to simulate a soft storey condition; both tests were performed for the high level inten-

sity (1.5).

• Final cyclic test which was carried out quasi-statically, again on the bare structure, for three

levels of increasing top-displacement, in order to assess the structure behaviour at a near-

failure stage.

All these tests were performed in the direction indicated in Figure 5.3, where the positive

direction convention for displacements is also indicated. Between each test the structure was

unloaded to zero force, since the actuators had to be removed before a new testing phase.

Details are given in the next paragraphs for each testing phase. The main global results are pre-

sented and commented, after brief references to the test setup and the input accelerogram. The

following abbreviations are adopted to identify the tests: 0.4S7 (or low level) and 1.5S7 (or

high level), standing for the low and the high level seismic tests on the bare frame, uniform

and soft-storey for the two configurations of masonry infilled frame tests and Duct.3, Duct.5

and Duct.8 for the three phases of the final cyclic test (as detailed in 5.5.6).

5.5.1 PSD test setup

The kinematic degrees of freedom considered for the PSD tests consisted of four slab horizon-

tal displacements in the testing direction, which are sufficient to adequately describe the struc-

ture behaviour for lateral actions.

No rotational d.o.f. in the floor plan, nor translational ones in the transverse direction were

considered in the equations of motion. Therefore, these d.o.f. had to be restrained in the exper-

imental setup, which has been achieved by the following:

• the storey forces necessary to drive the structure were provided by means of a pair of digit-

Page 233: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 209

ally servo-controlled hydraulic actuators per floor (attached to the external frames) which

imposed the same displacement and thus restrained the floor rotation;

• in order to prevent translation in the transversal direction, another actuator was horizontally

placed in the third floor, parallel to the reaction wall, to which a zero displacement condi-

tion was imposed during the test.

Picture 5.1 shows a lateral view of the structure including the main actuator setup and the steel

reference frames for displacement measurements.

Picture 5.1 Lateral view of the structure with actuator and reference frame setup

Forces were measured by load cells placed at the end of each actuator piston rod. Piston con-

trol was done by an optical digital encoding device, through which storey-slab displacements

were measured relative to the steel reference frames mounted on the testing platform.

In correspondence with the four d.o.f., a lumped mass matrix was considered. In order to

account for the additional dead loads and factorized live loads, supposed to act simultaneously

with seismic forces, additional masses were attached to each floor (namely 24.3 ton for the

first three storeys and 26.1 ton for the top storey) and included in the mass matrix.

Page 234: Seismic Assessment of Reinforced Concrete Frame Structures

210 Chapter 5

Page 235: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 211

Since only four d.o.f. are present, not very stringent constraints are imposed to the integration

time step. Therefore the central difference method was adopted for PSD testing with a time

interval of 4 ms and 2 ms, respectively, for the bare frame and for the infilled frame tests.

The viscous damping factor was set to zero, since the energy in earthquake-like vibration is

essentially dissipated by hysteresis which comes automatically included in the response of the

physically tested specimen.

Details on the whole instrumentation setup can be found in Negro et al. (1994), but the main

types of measurements are briefly referred herein for completeness:

• Displacements and restoring forces of the four storeys.

• Total rotations within the potential plastic hinge zones (critical zones), at the ends of all the

beams and at the base level of columns; in beams the measurement base length was 450 mm

(the full depth of beams) and in columns two base lengths were adopted, namely 225 mm

and 450 mm.

• Distribution of rotations inside the critical zones of the second storey beams, by means of

measurements on three base lengths: 60 mm, 225 mm and 450 mm (measured from the side

of columns).

• Deformations of the top face of the slab in the testing direction, along the transversal beams

of the second storey, in order to estimate the slab contribution to the stiffness and strength of

the frames.

5.5.2 The input accelerogram

A set of ten artificial accelerograms were generated by Pinto and Pegon (1991), fitting the EC8

response spectrum for soil type B and a damping factor of 5%, for which a peak ground accel-

eration of 0.30g and 10s of duration was prescribed. Out of these artificial signals, modulated

by waveforms identified from the 1976 Friuli earthquake, two were selected for preliminary

non-linear analyses carried out by several members of the EASML (Carvalho (1993), Negro et

al. (1994)). From such analyses, one signal was indicated by all members as the most demand-

ing and, therefore, it was chosen as the reference signal for the experimental seismic tests. Des-

ignated by S7, the adopted accelerogram is shown in Figure 5.4 together with its elastic

response spectrum, which is compared with the reference response spectrum of EC8 (soil type

B and 5% damping).

∆t

Page 236: Seismic Assessment of Reinforced Concrete Frame Structures

212 Chapter 5

Figure 5.4 Input accelerogram (Friuli-like) and elastic response spectra

For both experimental and numerical analyses described in the present and the following chap-

ters, two levels of seismic action intensity were considered:

• the low level, corresponding to 0.4 times the design acceleration (thus, 0.12g), was chosen

to approximately represent the seismic event within the serviceability state as established in

paragraph 4.3.2 of Part 1-2 of EC8;

• the high level, associated with a peak ground acceleration of 1.5 times the design one (thus,

0.45g), was intended to induce net inelastic effects on the structure.

It is worth stressing that the intensity factor was set up to scale the design peak ground acceler-

ation (0.30g); however, because the peak acceleration of the reference signal is actually 0.38g

(thus, about 27% higher than the design acceleration), the resulting peak accelerations for the

low and high level tests reach 0.15g and 0.57g, respectively. Therefore, if test results are to be

compared with design values (e.g. for base-shear), one has to bear in mind such discrepancy

between the design assumption and the actually adopted peak ground acceleration.

Time (s)

Ground acceleration[g]

Period (s)

Spectral acceleration [g]

0.0 0.4 0.8 1.2 1.6 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1

EC8 Spect.

2.0

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 -.40 -.32 -.24 -.16 -.08 .00 .08 .16 .24 .32 .40

a) Artificially generated accelerogram S7

b) Elastic response and EC8 spectra

Max.: 0.38

Page 237: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 213

5.5.3 Preliminary tests

The following different types of preliminary tests were performed: dynamic snap-back tests,

their PSD simulation and direct stiffness tests.

From the dynamic snap-back tests the main initial frequencies, mode shapes and damping fac-

tor were envisaged. They were performed by pulling the structure towards the reaction wall by

means of a steel bar designed to break at a pre-defined load which would not introduced signif-

icant cracking in the structure. Two of these tests were carried out: one by pulling the top sto-

rey and the other by applying the force on the third storey, so that the four main vibration

modes could be captured. The free vibration time histories of storey displacements allowed to

obtain frequencies from the corresponding power spectra and the average damping ratio was

estimated as 1.8%.

The dynamic snap-back tests were simulated by means of PSD testing; the main objective was

to assess the PSD implementation algorithm. A null damping factor was considered for the vis-

cous damping forces in these PSD simulations. The results compared reasonably well (Negro

et al. (1994)) and even helped to detect a fault in the acquisition system during the dynamic

testing. The time histories of storey displacements from the PSD simulation led to an average

damping ratio of 2.2%, slightly higher than the previous one.

The direct stiffness tests were performed by displacing each floor by a prescribed value (small

enough in order to avoid significant cracking) while the others remained fixed. The restoring

force measurements led to the estimation of the stiffness matrix condensed to the four degrees

of freedom (one per floor). Little asymmetry on the stiffness matrix was found, which is una-

voidable since some stiffness degradation always occurs, even if negligible, between each test-

ing step. Using an estimate of the lumped mass matrix and the experimentally obtained

stiffness matrix, the vibration frequencies were calculated (Negro et al. (1994)) and are sum-

marized in Table 5.3 along with those obtained from the snap-back tests.

Frequency values for the various tests compare well and the fact that values are decreasing

from the dynamic test to the stiffness one, can be related with the cracking progression and

possibly with the adopted mass estimate, which is used to obtain the stiffness test frequencies

but not in the dynamic test. The same reason might explain the difference of damping ratios

given above.

Page 238: Seismic Assessment of Reinforced Concrete Frame Structures

214 Chapter 5

It is noteworthy that the PSD simulation without viscous damping forces in the dynamic equi-

librium equations, led to energy dissipation similar to that of the dynamic test. It follows that

dissipation had to be due to the hysteretic restoring forces measured during the PSD simula-

tion, because viscous damping forces were not included. This comparison of dynamic with

PSD simulation of snap-back tests confirms the hysteretic nature of structural damping.

5.5.4 Seismic tests on the bare frame structure

After the preliminary tests the structure was subjected to seismic tests of two intensity levels

whose main results are presented and commented in this section. Results of both tests are

shown together for easier comparison between them (see Figures 5.5, 5.6 and 5.7); generally,

the left-most side stands for the low level test and the right-most side to the high level test.

Figure 5.5 includes the time histories of storey displacements, of relative inter-storey drifts

(difference of successive storey displacements divided by the inter-storey height) and of total

inter-storey shear (the total horizontal force immediately below each floor level). The storey

profiles of peak values of those time histories are also included in the same figure. The storey

shear-drift diagrams, relating the inter-storey shear with the relative inter-storey drift, are

shown in Figure 5.6. The curves of dissipated energy at each storey, which are equal to the area

enclosed by the shear-drift diagrams multiplied by the storey height, are depicted in Figure 5.7,

also including the base shear - top displacement diagrams.

5.5.4.1 Low level test

For the low level test, intended to correspond to a serviceability limit state, neither significant

damage was observed nor apparent yielding seems to have occurred. In spite of the non-availa-

bility of internal force-deformation relationships to check the local section behaviour, it can be

roughly assessed by the storey shear-drift diagrams shown in Figure 5.6-a).

Table 5.3 Frequencies (Hz) for all testing cases.

Mode DynamicSnap-Back

PSDSnap-Back

StiffnessTest

1 1.90 1.85 1.782 5.95 5.54 5.123 10.40 9.94 8.654 16.30 13.5 12.00

Page 239: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 215

Figu

re 5

.5Ti

me

hist

orie

s of

sto

rey

disp

lace

men

ts, r

elat

ive

inte

r-st

orey

drif

t, to

tal s

tore

y-sh

ear a

nd re

spec

tive

peak

val

ue p

rofil

es fo

r

Low

and

Hi g

h le

vel t

estsTi

me

(s)

DIS

PL.(m

)

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

-.1

0

-.0

8

-.0

6

-.0

4

-.0

2

.0

.02

.04

.06

.08

.10

Tim

e (s

)

DIS

PL.(m

)

-.2

5

-.2

0

-.1

5

-.1

0

-.0

5

.0

.05

.10

.15

.20

.25

DIS

PL. (

m)

.0

.05

.10

.15

.20

.25

1

2

3

4

Low

Hig

h

STO

REY

21

34

Stor

ey:

DRI

FT (%

)

-1.0

-

.8

-.6

-

.4

-.2

.0

.2

.4

.6

.8

1.0

D

RIFT

(%)

-2.5

-2

.0

-1.5

-1

.0

-.5

.0

.5

1

.0

1.5

2

.0

2.5

.0

.50

1

.0

1.5

0

2.0

2

.50

1

2

3

4

Test:

21

34

Stor

ey:

STO

REY

Tim

e (s

)

SHEA

R (k

N)

-6.0

-4

.8

-3.6

-2

.4

-1.2

.0

1.2

2

.4

3.6

4

.8

6.0

x1

.E2

Tim

e (s

)

SHEA

R (k

N)

-1.5

-1

.2

-.9

-

.6

-.3

.0

.3

.6

.9

1

.2

1.5

x1

.E3

SHEA

R (x

1.E3

kN

)

.0

.3

0

.6

0

.9

0

1.2

0

1.5

0

1

2

3

4 ST

ORE

Y

Tim

e (s

)

Tim

e (s

)

DRI

FT (%

)

.0

.8

1.6

2.

4 3

.2

4.0

4.8

5.

6 6

.4

7.2

8.0

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

a) T

ime

hist

orie

s - L

ow le

vel

c) T

ime

hist

orie

s - H

igh

leve

l b

) Pea

k va

lues

Page 240: Seismic Assessment of Reinforced Concrete Frame Structures

216 Chapter 5

Figure 5.6 Shear-drift diagrams at each storey, for Low and High level tests

DRIFT (%)

SHEAR (kN)

-1.0 -.8 -.6 -.4 -.2 .0 .2 .4 .6 .8 1.0 -6.0 -4.8 -3.6 -2.4 -1.2 .0 1.2 2.4 3.6 4.8 6.0 x1.E2

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5 -1.5 -1.2 -.9 -.6 -.3 .0 .3 .6 .9 1.2 1.5

DRIFT (%)

SHEAR (kN) x1.E3

DRIFT (%)

SHEAR (kN)

-1.0 -.8 -.6 -.4 -.2 .0 .2 .4 .6 .8 1.0 -6.0 -4.8 -3.6 -2.4 -1.2 .0 1.2 2.4 3.6 4.8 6.0 x1.E2

DRIFT (%)

SHEAR (kN)

-1.0 -.8 -.6 -.4 -.2 .0 .2 .4 .6 .8 1.0 -6.0 -4.8 -3.6 -2.4 -1.2 .0 1.2 2.4 3.6 4.8 6.0 x1.E2

DRIFT (%)

SHEAR (kN)

-1.0 -.8 -.6 -.4 -.2 .0 .2 .4 .6 .8 1.0 -6.0 -4.8 -3.6 -2.4 -1.2 .0 1.2 2.4 3.6 4.8 6.0 x1.E2

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5 -1.5 -1.2 -.9 -.6 -.3 .0 .3 .6 .9 1.2 1.5

DRIFT (%)

SHEAR (kN) x1.E3

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5 -1.5 -1.2 -.9 -.6 -.3 .0 .3 .6 .9 1.2 1.5

DRIFT (%)

SHEAR (kN) x1.E3

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5 -1.5 -1.2 -.9 -.6 -.3 .0 .3 .6 .9 1.2 1.5

DRIFT (%)

SHEAR (kN) x1.E3

a) Low level b) High level

Storey4

Storey3

Storey1

Storey2

Low levelenvelope

Page 241: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 217

Figu

re 5

.7Ti

me

hist

orie

s of

dis

sipa

ted

ener

gy a

nd b

ase

shea

r - to

p di

spla

cem

ent d

iagr

ams

for L

ow a

nd H

igh

leve

l tes

ts

Tim

e (s

)

ENER

GY

(kJ)

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

3.5

7.0

10.

5

14.

0

17.

5

21.

0

24.

5

28.

0

31.

5

35.

0

Tim

e (s

)

ENER

GY

(kJ)

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

.3

5

.7

0

1.0

5

1.4

0

1.7

5

2.1

0

2.4

5

2.8

0

3.1

5

3.5

0

x1.E

2

STO

REY-

1

STO

REY-

2

STO

REY-

3

STO

REY-

4

TOTA

L

Low

leve

lH

igh

leve

l

TOP

DIS

PL.(m

)

BASE

SH

EAR

(kN

)

-.1

0 -

.08

-.0

6 -

.04

-.0

2

.0

.02

.0

4

.06

.0

8

.10

-6

.0

-4.8

-3

.6

-2.4

-1

.2

.0

1

.2

2.4

3

.6

4.8

6

.0

x1.E

2

TOP

DIS

PL.(m

)

BASE

SH

EAR

(kN)

-.2

5 -

.20

-.1

5 -

.10

-.0

5

.0

.05

.1

0

.15

.2

0

.25

-1

.5

-1.2

-

.9

-.6

-

.3

.0

.3

.6

.9

1.2

1

.5

x1.E

3

a) E

nerg

y di

agra

ms

b) B

ase

shea

r -

diag

ram

s

top

disp

lace

men

t

Low

leve

len

velo

pe

Page 242: Seismic Assessment of Reinforced Concrete Frame Structures

218 Chapter 5

For the low level test it can be seen that a major stiffness variation occurs in those diagrams,

particularly for the storeys 1, 2 and 3, corresponding to the extensive cracking produced by the

first significant displacement peak occurred around 2.2 seconds. A clear stiffness drop is

induced towards that peak and temporary large residual displacements took place after unload-

ing, which were due to member sections only partially cracked for one bending direction. After

a significant load reversal, these sections also cracked in the other bending direction and the

global section stiffness was reduced to the fully-cracked one; further unloading followed this

cracked stiffness and led to low residual deformations.

The early stage of damage corresponding to the low level test can be confirmed by comparing

the shear-drift diagrams between both level tests. For this purpose the envelopes of the low

level test diagrams are overprinted with thicker line in the corresponding high level test dia-

grams, in order to highlight the different global stiffness between the two tests. Additionally,

the amount of dissipated energy, also related to the damage state of the structure, is much lower

in the low level test, as can be qualitatively seen by the different size of shear-drift loops and

quantitatively checked in Figure 5.7-a). Note that the total final dissipated energy in the low

level test is at least one order of magnitude lower than the high level one.

Although not experimentally measured, the first mode frequency after the low level test can be

estimated from the last displacement cycles as 1.27 Hz. Figure 5.5-b) shows that the top dis-

placement reached the peak value of 3.7 cm (i.e. a total drift of 0.30%), the maximum inter-

story drift was 0.36% in the second storey and the peak base-shear was 583 kN.

Since these results arise from a seismic input approximately corresponding to the serviceability

state, it is apparent that the structure verifies the inter-storey drift requirement associated with

the EC8 serviceability limit state (paragraph 4.3.2 of Part 1-2) which is 0.4% for structures

“having non-structural elements of brittle materials attached to the structure” and 0.6% in case

such elements do not interfere with structural deformations.

Note that the maximum inter-storey drift occurs at the second storey due to the frame type

deformed shape. The first storey columns are restrained at the bottom by the foundation, while

the displacements of second storey columns mainly depend on the stiffness of the joints and

the framing beams. Thus, the more flexible boundary conditions at the second storey columns

lead to higher inter-storey drift, even for a shear force lower than for the first storey.

Page 243: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 219

From the overall inspection of the structure after the test, no significant permanent cracks were

visible apart from those detected before the test and possibly due to shrinkage effects; this

aspect sustains the small residual deformations as stated above.

5.5.4.2 High level test

The reference accelerogram was scaled by a 1.5 factor for the high level test, which was

thought to apparently damage the structure. Instead, this was not the case, since not very exten-

sive and permanent damage could be found from the post-test inspection.

However, during the test, clearly visible cracks opened for the maxima deformed shapes of the

structure, in the critical regions (member end zones) of the three first storey beams and of sev-

eral columns (mostly of the first storey), and in the two first storey beam-column joints.

The cracking pattern of beams consisted of one major crack at the beam-column interface,

which remained permanently open after the test, and several other cracks, less open and with

spacing increasing towards the mid-span. These cracks were inclined, due to internal shear

forces, with the inclination angle reducing for increasing distance from the end sections. In the

columns, cracks appeared more visible at the base level, with a pattern similar to that of beams.

In the beam-column joints of the first and second storey diagonal cracking patterns clearly

developed, showing that a diagonal strut mechanism was activated to transfer forces across the

joint.

Note that a very low dissipation capacity of the structure can be detected in the shear-drift dia-

grams shown in Figure 5.6-b). A detailed analysis of such diagrams showed that some dissipa-

tion effectively occurred until the maximum drift in each direction was reached (approximately

between 3 s and 4 s), after which the cycle diagrams became very pinched for both loading

directions. This aspect is illustrated in Figure 5.8 where, for the second storey, the shear-drift

diagram was split into two parts: one until 4 s of the test duration and another from that instant

till the end. Note that this drop in dissipation capacity is also apparent from the dissipated

energy curves shown in Figure 5.7-a), where the energy increase rate clearly reduces after 4 s.

Development of permanent major cracks at beam-column interfaces suggests that yielding of

rebars took place locally and the steel-concrete bond inside the joint core might have been seri-

ously damaged, leading to significant bar-slippage. This contributes to a stiffness reduction

Page 244: Seismic Assessment of Reinforced Concrete Frame Structures

220 Chapter 5

while cracks are totally open across the whole section depth, during the unloading-reloading

process; after crack closure, internal forces can be transferred by a main diagonal strut, for

which bond behaviour is not relevant, and stiffness increases again. As a result, force-displace-

ment diagrams show the pinched shape apparent in Figure 5.8-b).

Figure 5.8 Detail of the 2nd storey experimental shear-drift diagram for the high level test

Besides the described cracking pattern and the local yielding of bars at the beam-column inter-

faces, no other apparent signs of serious damage could be seen in the structure, namely spalling

of concrete cover or buckling of reinforcement.

By means of a stiffness test performed after the high level test, new vibration modes were com-

puted leading to the following frequencies (Hz): 0.82, 2.79, 5.19 and 7.34. The mode shapes

did not differ too much from the virgin structure ones, but the drop of frequency values is quite

significant. Note that the fundamental frequency became less than 50% of the initial value,

which means a reduction of the corresponding modal stiffness to about 20% of the value before

the low level test. This fact sustains the occurrence of non-linear phenomena clearly beyond

the cracking phase.

The peak value of top displacement was 21.2 cm (thus, a total drift of 1.70%), while the inter-

storey drift and base shear were 2.41% and 1435 kN, respectively. The maximum inter-storey

drift still occurred at the second level and, although not meaningful as a limit state verification,

it is worth noting that it is far beyond the EC8 thresholds for the serviceability limit state.

From the rotation measurements in the critical zones, the spatial distribution of the correspond-

ing peak values was plotted as shown in Figure 5.9, where the maximum rotation obtained is

14.9 mRad, while for the low level test it was 2.51 mRad.

-8.0 0.0 8.0 -1.5

-.9

.0

.9

1.5 x1.E3

DRIFT (cm)

-8.0 0.0 8.0 -1.5

-.9

.0

.9

1.5 x1.E3SHEAR (kN) SHEAR (kN)

DRIFT (cm)

a) From 0 to 4 s b) After 4 s

Page 245: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 221

Figure 5.9 Spatial distribution of rotations in the critical zones, for the high level test.

A good and almost uniform distribution of rotations was obtained among the beams of the

three first storeys, whilst the top storey beams show little contribution to the global deforma-

tion as can be confirmed in the dissipated energy curves of Figure 5.7. Note that, according to

5.5.1, for the critical regions of the second storey, rotations are split into two parts correspond-

ing to the two half lengths of each region.

Although useful for assessing the overall distribution, the results of Figure 5.9 do not help to

identify the mechanism of deformation developed, due to the lack of rotation measurements

for the columns. Since these measurements were made only at the expected plastic hinge loca-

tions, namely beam and base column ends, it is difficult to check, only on the basis of these

results, the effective formation of a weak beam - strong column mechanism.

5.5.5 Infilled frame seismic tests

Aiming at the study of the influence of infill panels on the behaviour of frame structures, two

pseudo-dynamic tests were performed with different infill lay-out. After the testing phase on

the bare structure, the external frames were infilled with light unreinforced masonry, first in a

regular configuration (uniformly infilled) and then in an irregular one, in which only the three

upper floors where infilled, simulating a structure with a “soft-storey”.

No repair was carried out before these tests because the damage induced by the previous ones

appeared to be relatively low. On the other hand, only high level intensity tests were per-

formed, because the fragile behaviour of mortar was not compatible with a low level test; the

same input signal was used as for the 1.5S7 test.

Experimental Rotations (mRad) - Max. = 14.9

Internal Frame External Frame

Page 246: Seismic Assessment of Reinforced Concrete Frame Structures

222 Chapter 5

The infill panels were made with typical materials, namely hollow ceramic bricks having the

dimensions of 245 x 112 x 190 (h) mm with vertical holes and mortar with average compres-

sive strength of 5 MPa. Compressive tests were performed on the blocks, parallel and orthogo-

nal to the holes, and small panel specimens were also tested to obtain the mechanical

properties of the masonry in three directions: the strongest one (parallel to the holes), the

weakest one (orthogonal to the first) and the diagonal one. From these tests, the compressive

strengths and the Young modulus in the perpendicular directions, as well as the shear modulus,

can be obtained. Details about these values can be found in Negro et al. (1995) and Combes-

cure (1996).

The infilled frame test results are briefly presented and commented in the following paragraphs

and, as before, results for the regular and the irregular configuration are displayed side by side,

the leftmost corresponding to the uniform configuration and the rightmost to the soft-storey

one. Time histories of storey displacements, relative inter-storey drifts and total inter-storey

shear are shown in Figure 5.10, which also includes the corresponding storey profiles of peak

values. The storey shear-drift diagrams are depicted in Figure 5.11.

5.5.5.1 Uniformly infilled configuration

By means of a new stiffness measurement before the PSD test, the vibration frequencies were

updated for this infilled configuration. The first mode frequency increased from 0.82 Hz, after

the bare frame tests, to 3.34 Hz as a consequence of the stiffening effect of the infill panels.

From Figure 5.10 the maximum values of top displacement, inter-storey drift and inter-storey

shear can be read as 8.0 cm, 1.12% and 2083 kN, respectively. The comparison with the corre-

sponding values for the 1.5S7 test highlights the increased stiffness, since the peak value of top

displacement reduced to less than 40% and the drift reduced to about 45%. Note that the inter-

storey drift peak value occurred in the first floor, whereas for the bare frame tests it was found

in the second floor.

The maximum base shear, as an indicator of the structure global strength, increases about 45%,

which, in spite of the important strength degradation, still can be considered as a sign of wor-

thy strength enhancement. Actually, from the comparison of shear-drift diagrams, the same

amount of drift leads to shear values for the infilled structure significantly higher than for the

bare structure.

Page 247: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 223

Figu

re 5

.10

Tim

e hi

stor

ies

of s

tore

y di

spla

cem

ents

, rel

ativ

e in

ter-

stor

ey d

rift,

tota

l sto

rey-

shea

r and

resp

ectiv

e pe

ak v

alue

pro

files

for

both

con

figur

atio

ns o

f inf

illed

fram

e te

sts

Tim

e (s

)

DIS

PL.(m

)

Tim

e (s

)

DIS

PL.(m

)

DIS

PL. (

m)

1

2

3

4

Uni

form

Soft-

stor

ey

STO

REY

21

34

Stor

ey:

DRI

FT (%

)D

RIFT

(%)

1

2

3

4 Te

st:2

13

4St

orey

:

STO

REY

Tim

e (s

)

SHEA

R (k

N)

x1.E

3

Tim

e (s

)

SHEA

R (k

N)

x1.E

3

SHEA

R (x

1.E3

kN

)

1

2

3

4 ST

ORE

YTi

me

(s)

Ti

me

(s)

D

RIFT

(%)

-.2

0

-.1

6

-.1

2

-.0

8

-.0

4

.0

.04

.08

.12

.16

.20

.0

.0

2 .0

4 .0

6 .0

8 .1

0 .1

2 .1

4 .1

6 .1

8 .2

0

-4.0

-3

.2

-2.4

-1.6

-

.8

.0

.8

1.6

2

.4

3.2

4

.0

.0

.4

.8

1.2

1.6

2.0

2.4

2.8

3.2

3.6

4.0

.0

.5

1.0

1

.5

2.0

2

.50

-2

.5

-2.0

-1

.5

-1.0

-

.5

.0

.5

1.0

1

.5

2.0

2

.5

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

-.2

0

-.1

6

-.1

2

-.0

8

-.0

4

.0

.04

.08

.12

.16

.20

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

-4.0

-3

.2

-2.4

-1.6

-

.8

.0

.8

1.6

2

.4

3.2

4

.0

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

-2.5

-2

.0

-1.5

-1

.0

-.5

.0

.5

1

.0

1.5

2

.0

2.5

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

a) T

ime

hist

orie

s - U

nifo

rm c

onfig

urat

ion

c) T

ime

hist

orie

s - S

oft-s

tore

y co

nfig

urat

ion

b) P

eak

valu

es

Page 248: Seismic Assessment of Reinforced Concrete Frame Structures

224 Chapter 5

Figure 5.11 Shear-drift diagrams at each storey for both configurations of infilled frame tests

DRIFT (%)

SHEAR (kN) x1.E3 a) Uniform b) Soft-storey

Storey4

Storey3

Storey1

Storey2

-4.0 -3.2 -2.4 -1.6 -.8 .0 .8 1.6 2.4 3.2 4.0 -2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

DRIFT (%)

SHEAR (kN) x1.E3

-4.0 -3.2 -2.4 -1.6 -.8 .0 .8 1.6 2.4 3.2 4.0 -2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

DRIFT (%)

SHEAR (kN) x1.E3

-4.0 -3.2 -2.4 -1.6 -.8 .0 .8 1.6 2.4 3.2 4.0 -2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

DRIFT (%)

SHEAR (kN) x1.E3

-4.0 -3.2 -2.4 -1.6 -.8 .0 .8 1.6 2.4 3.2 4.0 -2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

DRIFT (%)

SHEAR (kN) x1.E3

-4.0 -3.2 -2.4 -1.6 -.8 .0 .8 1.6 2.4 3.2 4.0 -2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

DRIFT (%)

SHEAR (kN) x1.E3

-4.0 -3.2 -2.4 -1.6 -.8 .0 .8 1.6 2.4 3.2 4.0 -2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

DRIFT (%)

SHEAR (kN) x1.E3

-4.0 -3.2 -2.4 -1.6 -.8 .0 .8 1.6 2.4 3.2 4.0 -2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

DRIFT (%)

SHEAR (kN) x1.E3

-4.0 -3.2 -2.4 -1.6 -.8 .0 .8 1.6 2.4 3.2 4.0 -2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

Page 249: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 225

For this test no rotation measurements were taken for beams and columns, but the panel distor-

tions were measured by means of a specific instrumentation scheme detailed in Negro et al.

(1995).

Damage was found mainly in the first and second floor panels, which is also reflected in the

shear-drift diagrams of Figure 5.11. Highly pinched diagrams were obtained due to the pro-

gressive deterioration of infills. However, the pinching effect was found from a very early

deformation stage, and also in the third and fourth storeys, due to the sliding in the masonry-

concrete interface, where cracks first appeared. Crushing of the masonry at the panel corners of

the first two levels was quite extensive and even led to the complete failure of one first level

panel where a horizontal slice at mid-height fell-off. At the third and fourth storeys no crushing

or significant cracking was observed.

5.5.5.2 Soft-storey configuration

For this configuration the panels were replaced with new ones but only in the three upper sto-

reys. This led to a “soft-storey” at the first level, for which no design provisions were taken as

would be required by EC8, namely a local increase of design forces and the detailing of the

entire column length as a critical region.

The new fundamental frequency turned out to be 1.67 Hz, therefore about half of the obtained

for the regular configuration. The peak values of top displacement, inter-storey drift and shear

were, respectively, 17.6 cm, 3.55% and 1690 kN.

As expected, the deformation was mainly concentrated at the first level, with an inter-storey

drift about three times higher than the second storey one. As a result, the part of the structure

above the first floor moved almost as a rigid body on the first level columns, which can be

understood from the results and was clearly seen during the test. This is confirmed by the low

damage in three upper floors, since only in the second floor some masonry crushing appeared.

Accordingly, the energy dissipation took place essentially at the first level, due to the deforma-

tion of the bare frame during the cycle with a peak drift not experienced before. After the peaks

in both loading directions, the shear-drift diagram (see Figure 5.11-b)) became very pinched,

due to anchorage pull-out at both ends of columns and beams, and significant strength reduc-

tion occurred due to concrete spalling clearly observed in the columns.

Page 250: Seismic Assessment of Reinforced Concrete Frame Structures

226 Chapter 5

It is worth noting the increase of maximum base-shear (over 15%) when compared with the

bare frame test results (both the seismic and the final cyclic ones). In the bare frame tests, as

well as in the soft-storey one, the base-shear is bounded by the first storey strength, which

depends on the developed deformation mechanism. For the bare frame tests, the inflection

points in the first storey columns are closer to the upper end sections and their resistance is not

fully engaged. On the contrary, for the soft-storey test such columns tend to have a more shear-

type deflected shape, the inflection points become closer to the mid-span and higher resisting

moments develop in the upper end sections; consequently an increased shear force is generated

in that storey, but for a much higher drift likely to correspond to excessive ductility demands.

5.5.6 Final cyclic tests

The near-failure behaviour of the bare frame structure was simulated by means of the final

cyclic tests, whose main scope was the assessment of ultimate displacement and energy dissi-

pation capacities and of the damage sensitivity to cyclic loading.

Due to the significant amount of damage suffered during the irregular infilled frame test,

mainly in the two first floors, the structure had to be repaired. The repair was made in the criti-

cal regions where concrete cover spalling had occurred and where permanently open cracks

were visible. This intervention consisted on removing the damaged cover and replacing it by a

new one made out of no shrinkage mortar reinforced with stainless steel fibres.

The final tests consisted on three sets of cycles of increasing top displacement amplitude as

shown in Figure 5.12, each set with three cycles at the same peak displacement as follows:

• The first set reached the maximum displacement of 21.0 cm, in both loading directions,

equal to the one of the high level test, and aimed, on one hand at approximately re-establish-

ing the conditions existing at the end of the 1.5S7 test and, on the other hand, at checking

the structural hysteretic behaviour for that level of displacement already experienced.

• The second set drove the structure to a top displacement of 35.0 cm, also in both loading

directions.

• In the last set, a maximum displacement of about 60.0 cm was imposed to the structure in

the negative direction (pushing the structure away from the reaction-wall), whilst in the

other direction only 35.0 cm were reached due non-symmetric actuator stroke.

Page 251: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 227

Figure 5.12 Time history of the imposed top displacement for the final cyclic tests

Each of these sets corresponded to a certain level of global ductility. Actually, from the base

shear - top displacement diagrams of the 0.4S7 and 1.5S7 tests, the yielding value of top dis-

placement was roughly estimated as 7 cm; thus, the first cycle set corresponds to a ductility

level of 3, the second one to a ductility of 5 and the last one attained the ductility of 8 in one

testing direction. Therefore, from now on these tests are identified by the corresponding ductil-

ity level, respectively Duct.3, Duct.5 and Duct.8, and, although they actually consisted in a

unique test, the results are split according to each ductility level; the corresponding total drifts

were 1.68%, 2.8% and 4.8%, respectively.

The driving signal was set as the top displacement according to Figure 5.12, but an inverted tri-

angular distribution of forces was actually prescribed to the actuators. The cyclic response is

illustrated in Figure 5.13 by means of the base shear - top displacement diagrams and the total

deformation energy curves for each ductility level testing phase. The base shear - top displace-

ment diagrams were preferred here because they are essentially similar to the storey shear-drift

curves. The peak value profiles of storey displacements, inter-storey drift and shear are

depicted in Figure 5.14.

For ductility level 3, a pinched force-displacement diagram was obtained as it had already

occurred for the post-peak response of the 1.5S7 and the soft-storey tests. Since displacements

did not exceed the peak values previously experienced, the deformation and strength mecha-

Step

DISPL. (m)

0 2700 5400 -.60

-.48

-.36

-.24

-.12

.0

.12

.24

.36

.48

.60

Duct. 3 Duct. 5 Duct. 8

Page 252: Seismic Assessment of Reinforced Concrete Frame Structures

228 Chapter 5

nisms remain essentially the same and are strongly influenced by the joint behaviour as already

explained. Therefore, the hysteresis loops still exhibit very low energy dissipation capacity,

with little strength deterioration; nevertheless, quite stable diagrams are obtained.

The maximum inter-storey drift was 2.24%, still occurring at the second storey, whilst the

maximum base shear was 1222 kN. The drift value agrees well with that of the 1.5S7 test, but

the base shear decreased, which, in spite of the repair, reflects the stiffness deterioration

induced by the soft-storey test.

The visual inspection of the specimen at the last peak for ductility level 3, showed little

spalling of concrete cover at the base section of first storey columns and diagonal cracking in

the central beam-column joints of the first and second storeys.

The progression towards ductility level 5, clearly introduced significant damage in the struc-

ture. The maximum inter-storey drift reached 3.9%, still at the second storey although very

close to that of the first storey (see Figure 5.14-b)), and the maximum base shear turned out to

be 1444 kN, therefore a similar value to that of the 1.5S7 test. The base shear - top displace-

ment diagram became much more dissipative, but quite apparent strength degradation occurred

for loops after the first. The observed behaviour for this testing phase was as follows:

• At the first positive displacement peak cracks of about 3-4 mm opened at the column base

sections and a global deformed shape of soft-storey type could be seen, showing the defor-

mation mechanism and the damage localization in the first storey due to the soft-storey test.

• The subsequent negative peak increased the crack opening at the column base sections to 5-

7 mm and showed wide cracks in beams and joints (about 3-5 mm); moreover, crushing of

concrete cover at the column bases started.

• In the following displacement peaks, some joints were fairly cracked (diagonal pattern as

shown in Picture 5.2), full depth cracks could be seen at beam-column interface sections of

the first storey and apparent spalling of concrete cover occurred, both in columns and in

some beam end zones which had not been repaired.

• In the first storey exterior joints adjacent to the shorter bay, the concrete cover spalled in the

external face of the column, i.e. in the transversal frame plane, as a result of the stress trans-

fer between the reinforcement and the concrete within the bent part of the anchorage length

(Paulay and Priestley (1992)); this is shown in Picture 5.3, actually referring to the structure

final stage but included here because this phenomenon started at ductility level 5.

Page 253: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 229

Figure 5.13 Base shear - top displacement diagrams and curves of total deformation energy for

the final cyclic tests

Top Disp [m]

Base Shear (kN)

-.60 -.48 -.36 -.24 -.12 .0 .12 .24 .36 .48 .60 -1.50

-1.20

-.90

-.60

-.30

.0

.30

.60

.90

1.20

1.50 x1.E3

b) Energy

Step

ENERGY (kJ) x1.E3

ENERGY (kJ) x1.E3

ENERGY (kJ)

0 900 1800 .0

.15

.30

.45

.60

.75

.90

1.05

1.20

1.35

1.50 x1.E3

0 900 1800 .0

.15

.30

.45

.60

.75

.90

1.05

1.20

1.35

1.50

0 900 1800 .0

.15

.30

.45

.60

.75

.90

1.05

1.20

1.35

1.50

Step

Step

Storey 1

Storey 2

Storey 3

Storey 4

TOTAL

Duct. 3

Duct. 5

Duct. 8

Top Disp [m]

Base Shear (kN)

-.60 -.48 -.36 -.24 -.12 .0 .12 .24 .36 .48 .60 -1.50

-1.20

-.90

-.60

-.30

.0

.30

.60

.90

1.20

1.50 x1.E3

Top Disp [m]

Base Shear (kN)

-.60 -.48 -.36 -.24 -.12 .0 .12 .24 .36 .48 .60 -1.50

-1.20

-.90

-.60

-.30

.0

.30

.60

.90

1.20

1.50 x1.E3

a) Base shear - Top displacement

Page 254: Seismic Assessment of Reinforced Concrete Frame Structures

230 Chapter 5

Figure 5.14 Storey profiles of peak values of displacement, inter-storey drift and inter-storey

shear for the final cyclic tests

DISPL. (m)

STOREY

.0 .06 .12 .18 .24 .30 .36 .42 .48 .54 .60

1

2

3

4

Duct. 3

Duct. 5

Duct. 8

DRIFT (%) .0 .75 1.50 2.25 3.00 3.75 4.50 5.25 6.00 6.75 7.50

SHEAR (x1.E3 kN) .0 .15 .30 .45 .60 .75 .90 1.05 1.20 1.35 1.50

Duct. 3

Duct. 5

Duct. 8

STOREY

1

2

3

4

STOREY

1

2

3

4

Duct. 3

Duct. 5

Duct. 8

a) Maximum StoreyDISPLACEMENTS

b) Maximum Inter-StoreyDRIFT

c) Maximum StoreySHEAR

Page 255: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 231

Picture 5.2 Diagonal cracking pattern in the beam-column joint (final test - Ductility 5)

Picture 5.3 Cracking pattern in a 1st storey joint in the external face of the column

Page 256: Seismic Assessment of Reinforced Concrete Frame Structures

232 Chapter 5

Page 257: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 233

The testing phase for ductility level 8, proceeded with heavy damage in the structure. In several

critical regions of the first and second storey, concrete crushing and subsequent cover spalling

was apparent, followed by stirrup failure and buckling of rebars, some of which ruptured in the

subsequent cycles. This is evidenced in Picture 5.4-a) showing reinforcement instability at one

first storey beam, of which at least one bar was visibly ruptured at the end of the test (see Pic-

ture 5.4-b)). Similarly, Picture 5.4-c) shows the rupture of a 20 mm bar of the central column at

the base section.

The extent of slab participation for such a high deformation level is also apparent in Picture

5.4-a) showing slab cracks developed parallel to the transverse beams along their total length.

The overall cracking pattern at an exterior first storey beam is illustrated in Picture 5.5-a).

Higher cracking density and more inclined cracks are found in the shorter span (on the left

side) than in the larger one, confirming the predominancy of lateral load over the vertical load

effects in the shorter span. A detailed view of cracking near the leftmost column is shown in

Picture 5.5-b), where a permanent full depth crack can be seen at the beam-column interface.

The base shear - top displacement diagram was still found very dissipative although with even

more strength degradation, which is confirmed by the maximum base-shear of 1425 kN,

slightly lower than that for the Duct.5 and reached for a higher displacement.

The maximum inter-storey drift increased up to 7.2%, approximately uniform in the first and

second storeys, corresponding to the structure deformed shape shown in Picture 5.6 taken at

the last top displacement peak of 60 cm. From the observed damage and the strength deteriora-

tion evidenced in the force-displacement diagrams, it can be concluded that a near-failure stage

was actually reached.

During the evolution from the Duct. 3 to the Duct. 8 phases, the first storey drift increased and

the difference of drift values between the first and the second storeys was progressively

reduced, so that, at the end, an almost uniform drift was obtained in these two storeys. This

was due, not only to the damaged beam-column joint at the first storey, but also to the progres-

sive deterioration of the critical region at the column-foundation interface where plastic hinges

actually developed.

Page 258: Seismic Assessment of Reinforced Concrete Frame Structures

234 Chapter 5

Page 259: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 235

Pict

ure

5.4

B

uckl

ing

and

rupt

ure

of b

eam

and

col

umn

rein

forc

emen

t bar

s (fi

nal t

est -

Duc

t. 8

)

a)

b)

c)

Page 260: Seismic Assessment of Reinforced Concrete Frame Structures

236 Chapter 5

Page 261: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 237

Picture 5.5 Final overall cracking pattern of a 1st storey exterior beam

a) Final cracking pattern of beam

b) Detail of crack in the beam-column interface

Page 262: Seismic Assessment of Reinforced Concrete Frame Structures

238 Chapter 5

Page 263: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 239

Picture 5.6 General view of deformed shape at the last peak of 60 cm top displacement for the final test - Duct.8

Page 264: Seismic Assessment of Reinforced Concrete Frame Structures

240 Chapter 5

Page 265: Seismic Assessment of Reinforced Concrete Frame Structures

THE 4-STOREY FULL-SCALE BUILDING TESTED AT ELSA 241

Note that force-displacement diagrams for ductility levels of 5 and 8 are considerably more

dissipative than those of the previous tests, not only because larger drifts are reached inducing

significant residual deformations, but also due to the lower pinching exhibited by the diagrams.

This fact is thought to be related with a lower influence of the joint behaviour and rebar slip-

page on the global deformation, and can be explained as follows:

• If a joint is severely damaged, the restraining effect of the framing beams on the columns is

drastically reduced, which means that columns behave more independently of beams; there-

fore, higher drifts are obtained, inducing column rotations larger than those on the beams

due to the non-effectiveness of force transfer across the joint.

• The pinching effect is less apparent in the columns than in the beams, on one hand due to

section symmetry and, on the other hand, because they are less prone to anchorage slippage

due to the beneficial effect of the axial force; consequently, if the behaviour becomes more

controlled by the columns, at least in the two first storeys, it follows that a less marked

pinching effect appears in the shear-drift diagrams.

5.6 Concluding remarks

In the preceding sections the results of the experimental testing of the four-storey reinforced

concrete building have been presented.

Reduced scale tests on cantilever specimens seemed to sustain the use of B500S Tempcore

steel within the EC8 requirements for ductile design of earthquake resistant structures,

although problems related with anchorage of rebars need to be further investigated.

Full-scale testing has started with snap-back (free-vibration) and stiffness tests to obtain initial

vibration frequencies (1.8 Hz for the fundamental one) and viscous damping factor (1.8%) of

the structure.

Pseudo-dynamic tests were performed in the bare frame structure, first for low level intensity

and then for high intensity earthquake. The former just led to cracking throughout the three

first storeys, responsible for a clear stiffness drop, but no evidence of yielding could be found.

In turn, for the high level intensity, further cracking developed, mainly along diagonals of

beam-column joints and locally at the beam-column interfaces where yielding of reinforce-

ment has occurred; besides a significant frequency (and stiffness) drop, a clear pinching effect

Page 266: Seismic Assessment of Reinforced Concrete Frame Structures

242 Chapter 5

was observed in force-deformation response diagrams, related to bar-slippage inside the joints

and responsible for rather low dissipation after the response peaks. However, an almost uni-

form distribution of rotations was obtained throughout the critical zones of beams (although

with little contribution of the top storey) and no serious permanent damage was found at the

end of these tests.

The structure with external infilled frames was tested for two configurations, viz a uniformly

infilled and a partially infilled, the latter simulating a “soft-storey” at the first level. Infills con-

sisted of typical unreinforced mansonry made out of hollow ceramic bricks and have contrib-

uted for a clear increase of the initial frequency and for some enhancement of the global

strength of the structure. Serious damage could be found in the masonry panels of the two first

storeys of the uniformly infilled configuration, reflected in a very clear pinching effect in the

response diagrams. The test on the “soft-storey” configuration led to visible damage in the first

storey, above which the upper floors moved almost as a rigid body, thus leading to energy dis-

sipation mainly concentrated in that storey. This test was found critical for the structure, since

no specific design provisions were included as required by EC8 to account for the presence of

infills and, particularly, of a “soft-storey”. Consequently, significant strength reduction

occurred due to concrete spalling observed mainly in the first storey columns.

After repairing of the damaged zones, a final cyclic test was quasi-statically performed again

in the bare frame structure for three sets of cycles at increasing top displacement amplitude,

whose final deformation stage reached a global ductility factor around 8.

The final test first produced some damage signs already obtained in previous tests and progres-

sively introduced heavier damage consisting of concrete crushing and cover spalling, which

subsequently led to stirrup failure and buckling of rebars. The clear strength degradation and

the high inter-storey drifts (around 7%) in the two first storeys, exhibited for the stage at global

ductility 8 corresponded to a near-failure state of the structure at the end of the final test.

Page 267: Seismic Assessment of Reinforced Concrete Frame Structures

Chapter 6

ANALYSIS OF THE 4-STOREY FULL-

SCALE BUILDING

6.1 Introduction

The experimental campaign on the four-storey reinforced concrete building described in the

previous chapter, provides an excellent means of calibration and assessment of numerical mod-

els for global seismic behaviour simulation.

Indeed, the availability of different tests for distinct stages of the structure behaviour (pre-

yielding and post-yielding for increasing ductility levels up to failure) and the use of the PSD

method (which is an essentially static test from the strictly experimental point of view) have

rendered the outcome of these tests rather suitable for comparison with numerical simulations

throughout various behaviour stages, both in quasi-static or dynamic conditions.

Particularly, if only the global behaviour is sought (say at the storey, or even at the element

level) the experimental tests have provided the fundamental output to be compared against the

response obtained by global element models as the one developed in the present study. How-

ever, notwithstanding the valuable information obtained, some additional local measures

would have been desirable to assess the section behaviour in critical zones (beam and base-col-

umn plastic regions, beam-column joints, etc.) from which a better insight of the global ele-

ment behaviour could be have been obtained. Unfortunately this was not possible and the

experimental-analytical comparisons are mostly performed at the storey level, reflecting the

behaviour of the relevant elements.

In this context, the main scope of this chapter is the assessment of the flexibility element model

Page 268: Seismic Assessment of Reinforced Concrete Frame Structures

244 Chapter 6

(as presented in Chapter 3), along with the auxiliary procedure to characterize the local section

behaviour (described in Chapter 4), for the simulation of the global behaviour of members

integrated in a complete frame structure, by comparing the numerical response against the glo-

bal structural results from the full-scale tests.

The whole set of data and modelling assumptions for numerical simulations are described in

6.2, covering relevant aspects for the frame structure modelling and discretization, the section

behaviour, the structural damping and the mansonry infill modelling.

A key issue in the seismic assessment of structures is the quantification of damage induced by

earthquake action, in order to determine the closeness of the structure state in relation to pre-

defined limit states. To this end, a brief review of available proposals of damage indices is

included in 6.3 and the damage index adopted in the present work context is further detailed,

particularly in what concerns the involved parameter quantification.

The major concern of this chapter, i.e., the numerical simulation of the experimental campaign,

is extensively described in 6.4. A broad set of numerical analyses is included, covering static

pushover analyses (to roughly check the monotonic structural behaviour), the static and

dynamic simulations of seismic tests (on bare and infilled structure configurations) and the

analyses for the quasi-static final cyclic test. Additionally, some static and dynamic calcula-

tions were performed also with a traditional model of fixed length plastic hinges, in order to

compare with the flexibility simulations. Finally, the main conclusions of this chapter are sum-

marized in 6.5, with particular emphasis on the numerical simulations of the four-storey build-

ing response.

6.2 Modelling assumptions and data

6.2.1 Structure modelling

6.2.1.1 Discretization

For the numerical analyses presented in this work, the structure was modelled in the testing

direction by means of a planar frame association. Due to symmetry conditions, only two

frames were considered: one to simulate the internal frame and another, with double stiffness,

vertical load and mass properties, to simulate the two external frames.

Page 269: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 245

Each beam and column was discretized by one flexibility global element as described in Chap-

ter 3, resulting in a very simple mesh of 40 elements and 30 nodes illustrated in Figure 6.1.

Rigid lengths were considered in the beams, measured between the column axes and the beam-

column interfaces. According to the capabilities of the flexibility element, distinct material and

geometrical properties were assigned where required to the left and right element nodes and

assumed uniform between each end section and the mid-span section.

Figure 6.1 Mesh for the structural analysis using flexibility global elements (dimensions in m)

The rigid floor diaphragm assumption was accomplished by imposing the same horizontal dis-

placement to all nodes at the same floor.

6.2.1.2 Collaborating slab width

The participation of slabs in the strength and stiffness of beams is a rather difficult issue to

quantify due to the wide range of involved factors. Usually it is measured by an equivalent

width , assumed fully collaborating with the beam deformation, which is taken uniform

along the full length of the beam or portions of it. In the following, refers only to the slab

contribution, meaning that the beam web width is not included, and the effective slab

width is given by .

A major difficulty in estimating relies on the concept itself: the equivalent slab width for

quantifying the structural stiffness is different from the one involved in beam strength calcula-

tions (Tjebbes (1994)). In the stiffness case, the slab participation along the full length of the

beam is relevant, and it is meaningful, for instance, when a good estimation of structural fre-

quencies is envisaged. For strength computation, the slab contribution must be evaluated at the

Internal Frame External Frame

X

Z

3.00

3.50

3.00

3.00

6.004.006.004.00Storey

4

3

2

1

Level

beq( )

beq

bw( )

beff bw beq+=

beq

Page 270: Seismic Assessment of Reinforced Concrete Frame Structures

246 Chapter 6

locations where maximum bending moments are expected and, as explained below, it is an

important issue for an adequate assessment of the structural mechanisms of failure. Therefore,

the modelling of slab participation in a given structure by means of equivalent slab width, aim-

ing at a good representation of both stiffness and strength, appears as a quite hard task to be

accomplished.

In particular, the slab contribution for the strength case is twofold:

• When the beam is bent with the top reinforcement bars in tension, the slab rebars are also

tensioned and act as an extra flexural reinforcement for the beam, up to a transverse dis-

tance from the beam that depends on the loading intensity. The so-obtained extra flexural

strength must be taken into account in order to correctly anticipate the failure mechanism

and to be coherent with the capacity design philosophy subjacent to modern design codes

(Eurocode 8 (1994)). If a strong column - weak beam mechanism is expected, the non-con-

sideration of such extra strength can indeed prevent that mechanism to develop and may

lead to a very different, and not desired, behaviour for the structural numerical model

(Paulay and Priestley (1992)).

• For compressive strain developing on the top face of the beam, the corresponding stresses

can spread along the slab and decrease the neutral axis depth. As explained for the reduced

scale tests in 5.4, this can lead to higher tensile strains in bottom bars and increase the ten-

dency to buckle in subsequent cycles. The inherent strength deterioration in cyclic behav-

iour can be analytically captured only if the collaborating slab width is included and,

obviously, if the adopted model is able to treat such a phenomenon.

From these two points, the first appears to be the most important, but the related uncertainties

are substantial. Above all, the fact that the loading intensity induces an increase of ,

because larger plastic hinge rotations near the column faces will engage more slab steel placed

farther away from the column, renders more difficult the structural modelling for analysis of

increasing earthquake intensity levels. Moreover, the effectiveness of slab steel also depends

on the existence of transversal beams; generally, the presence of such beams, implies a larger

slab participation.

Also, the zone where the slab contribution is to be assessed has influence on for strength

purposes. For plastic hinges next to external columns, the slab width is lower than for internal

plastic hinges, due to different mechanisms of membrane force transfer between the slab and

beq

beq

Page 271: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 247

the beams. Additionally, the value is also affected depending on whether the slab is com-

pressed or tensioned, which complicates even more the modelling for cyclic loading.

Due to all the uncertainties regarding the adequate assessment of the equivalent slab width and,

since the whole experimental testing programme on the structure under analysis involved sev-

eral loading intensities and distinct test conditions (for which the slab participation was cer-

tainly variable), it was decided to adopt a uniform estimate for over the whole structure.

In the preliminary numerical analyses reported in Carvalho (1993) and performed by the JRC

team prior to the tests, the values of were adopted according to an approximation proposed

in Park and Paulay (1975). The slab width was taken up to a distance of measured from

the beam face to each side, where is the slab thickness. Thus, for the present case the fol-

lowing effective slab width values were obtained:

• for the external frames:

• for the internal frames:

After the bare frame tests, an in depth analysis of the measurements for the assessment of the

collaborating slab width was carried out and reported in Tjebbes (1994). As already mentioned

in 5.5.1, these measurements were made in the second storey and time histories of deforma-

tions on the top face of the slab in the testing direction were obtained. The transversal profile

of strains on the slab could be traced along the internal and one external transversal beams,

from which the portion of slab participating in the beam deformation was estimated.

The obtained values of , for the displacements peaks of both the 0.4S7 and the 1.5S7 tests,

confirmed the distinct extent of slab collaboration for the low and the high level tests, as well

as the larger slab widths next to the transverse internal frame when compared with those adja-

cent to the external transverse beams. Moreover, the values of slab width more adequate for

stiffness estimation purposes, were, as expected, lower than those for strength analysis.

However, taking an average uniform value for in both bending directions it was found

(from experimental measurements) that the above adopted widths are about 20% underesti-

mated for the strength case, whilst for the stiffness case those values are overestimated by

around 10% along the external longitudinal frames and underestimated by about 20% along the

central frame. This shows that the adopted values provide a reasonable approximation of the

average slab width for the whole structure.

beq

beq

beq

4hs

hs

beff bw 4hs+ 0.90 m= =

beff bw 2 4hs( )+ 1.50 m= =

beq

beq

Page 272: Seismic Assessment of Reinforced Concrete Frame Structures

248 Chapter 6

In our opinion, it would be pointless to consider a more rigorous fitting of with the experi-

mentally obtained results, namely by considering different values according to the slab loca-

tion, the loading direction or intensity, since other relevant phenomena could not be taken into

account. Among these, the following are highlighted:

• The collaborating slab width depends on the deformation level and, particularly, on the

inter-storey drift; since it varies along the height, it follows that values for the top sto-

rey, for example, are not the same as those obtained for the second floor.

• Measurements were made by displacement transducers located only on the top face of the

slab, and the resulting deformations were considered, in the calculation of , as the aver-

age membrane strains in the slab; actually, the membrane strains are lower than the top face

ones, but there was no measurement to account for that difference, which is thought to be

significant since the slab thickness is 1/3 of the total beam depth.

From the explained reasons, it was decided to keep the estimates of equivalent slab width as

used in the preliminary analyses, i.e., with the values given above.

6.2.1.3 Mass and vertical static loads

The structural mass was quantified on the basis of an average value of the concrete unit weight

of 25 kN/m3. For the dead loads, the contributions of the 0.15 m thick reinforced concrete slab

and of the beam and column gross sections were taken into account. The additional loads (fin-

ishings and factorized live loads), as used in the test and already mentioned in 5.5.1, were

included in the total mass of each floor. Table 6.1 summarizes the mass contributions from

dead and additional loads, and the total mass considered per floor.

For the mass matrix computation, each floor mass was distributed over all the nodes belonging

to the floor, leading to a concentrated mass system with no rotational inertia.

Table 6.1 Floor mass values and vertical loads on beams

Mass (103 kg) Vertical loadsFloor Dead loads Additional Total (kN/m)

4 58.5 26.1 84.6 35.0

3 64.2 24.3 88.5 34.0

2 64.2 24.3 88.5 34.0

1 65.1 24.3 89.4 34.0

beq

beq

beq

Page 273: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 249

The vertical loads acting simultaneously with the seismic loads are due to the slab dead loads

and the beam self-weights, plus the factorized live loads. Due to symmetry conditions, the

same uniformly distributed vertical loads per unit length have been considered in the external

and internal frame beams, as also listed in Table 6.1. It is recognized however, that a more

refined quantification of the vertical load effects would require triangular distributions of load

per unit length.

6.2.2 Cross-section characteristics and material properties

6.2.2.1 Cross-sections

The reinforcement and cross-section details are extensively reported in Negro et al. (1994).

However, for the sake of completeness they are summarized in Figures 6.2, 6.3 and 6.4.

Figure 6.2 Typical beam and column cross-sections for both external and internal frames

For the beam cross-sections the top slab reinforcement mesh inside the slab width was consid-

ered part of the beam top reinforcement, whilst the mesh of the slab bottom face was neglected.

a) Internal frame

At

Ab

b) External frame

At

Ab

Asw

0.30 m

0.90 m1.50 m

0.30 m

0.45

m

0.15

m

# Q188 (DIN) = φ6//0.15

0.40 m

0.40

m

0.40 m

0.45

m

0.45 m0.40 m

a2

a1

a1

Beams

Columns

Lateral CentralLateral Central

Page 274: Seismic Assessment of Reinforced Concrete Frame Structures

250 Chapter 6

Figu

re 6

.3Sc

hem

atic

rein

forc

emen

t lay

out f

or th

e be

ams

a) In

tern

al fr

ame

3.5

2φ14

1φ12

3.5

2φ14

1φ12

5.2

2φ14

3φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

3φ12

5.2

5.2

3φ14

3φ12

5.2

3φ14

3φ12

5.2

3φ14

3φ12

3.5

3φ14

1φ12

3.5

3φ14

1φ12

3.5

3φ14

1φ12

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

3φ14

1φ12

3.5

5.2

5φ14

2φ12

5.2

5φ14

2φ12

5.2

5φ14

2φ12

3.5

5φ14

3.5

5φ14

3.5

5φ14

4φ14

3.5

4φ14

3.5

4φ14

3.5

4φ14

3.5

4φ14

3.5

4φ14

3.5

3.5

5φ14

2φ12

3.5

5φ14

2φ12

3.5

5φ14

2φ12

3.5

5φ14

1φ12

3.5

5φ14

1φ12

5.2

5φ14

2φ12

5φ14

5.2

4φ14

3.5

4φ14

3.5

4φ14

3.5

4φ14

3.5

4φ14

3.5

a 2 =

... (c

m)

a 1 =

... (c

m)

a 2 =

... (c

m)

a 1 =

... (c

m)

a 2 =

... (c

m)

a 1 =

... (c

m)

a 2 =

... (c

m)

a 1 =

... (c

m)

φ6//8

φ6//8

φ6//8

φ6//8

φ8//2

0φ8

//20

110

112.

511

2.5

110

Stirr

ups

φ(m

m)//

(cm

)

Dis

tanc

e fr

om

Col

umn

Cen

treLi

ne (c

m)

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

2φ14

1φ12

3.5

3.5

4φ14

3.5

4φ14

3.5

4φ14

3.5

4φ14

3.5

4φ14

3.5

4φ14

3φ14

1φ12

3.5

3φ14

1φ12

3.5

3φ14

1φ12

3.5

3φ14

1φ12

3.5

3φ14

1φ12

3.5

3φ14

1φ12

3.5

5.2

4φ14

2φ12

5.2

5φ14

5.2

5φ14

5.2

5φ14

5.2

4φ14

5.2

4φ14

2φ12

5φ14

5.2

4φ14

5.2

4φ14

5.2

4φ14

5.2

3φ14

5.2

3φ14

3.5

5.2

6φ14

5.2

6φ14

5.2

6φ14

5.2

6φ14

5.2

6φ14

5.2

6φ14

4φ14

2φ12

5.2

3φ14

2φ12

3.5

3φ14

2φ12

3.5

3φ14

2φ12

3.5

3φ14

2φ12

3.5

3φ14

2φ12

3.5

φ6//8

φ6//8

φ6//8

φ6//8

φ8//2

0φ8

//20

110

112.

511

2.5

110

b) E

xter

nal f

ram

e

Page 275: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 251

Figu

re 6

.4Sc

hem

atic

rein

forc

emen

t lay

out f

or th

e co

lum

ns

8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ20

4.0

8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ20

4.0

4φ16

+ 8

φ14

3.8 4φ16

+ 8

φ14

3.8 4φ16

+ 8

φ14

3.8 4φ16

+ 8

φ14

3.8 4φ16

+ 8

φ14

3.8 4φ16

+ 8

φ14

3.8 4φ16

+ 8

φ14

3.8 4φ16

+ 8

φ14

3.8 12φ1

63.

8 12φ1

63.

8 12φ1

63.

8 12φ2

04.

0

φ10/

/7.5

φ8//2

0

φ10/

/7.5

φ10/

/7.5

φ8//2

0

φ10/

/7.5

φ10/

/7.5

φ8//2

0

φ10/

/7.5

φ10/

/7.5

φ8//2

0

φ10/

/7.5

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

Stirr

ups

φ(m

m)//

(cm

)St

eel B

ars

a 1 =

a2 (

cm)

a) In

tern

al fr

ame

b) E

xter

nal f

ram

e

8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 4φ20

+ 4

φ16

4.0 4φ20

+ 4

φ16

4.0 4φ20

+ 4

φ16

4.0 4φ25

+ 4

φ16

4.25

8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 8φ16

3.8 4φ20

+ 4

φ16

4.0 4φ20

+ 4

φ16

4.0 4φ20

+ 4

φ16

4.0 4φ25

+ 4

φ16

4.25

4φ16

+ 8

φ14

3.8 4φ16

+ 8

φ14

3.8 12φ1

63.

8 12φ1

63.

8 12φ1

63.

8 12φ1

63.

8 12φ1

63.

8 12φ1

63.

8 12φ1

63.

8 12φ1

63.

8 12φ1

63.

8 12φ2

04.

0

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//2

0

φ10/

/10

φ10/

/10

φ8//1

0

φ10/

/10

φ10/

/10

φ8//1

0

φ10/

/10

φ10/

/10

φ8//1

0

φ10/

/10

φ10/

/10

φ8//1

0

φ10/

/10

φ10/

/10

φ8//1

0

φ10/

/10

φ10/

/10

φ8//1

0

φ10/

/10

φ10/

/10

φ8//1

0

φ10/

/10

φ10/

/10

φ8//1

0

φ10/

/10

Stirr

ups

φ(m

m)//

(cm

)St

eel B

ars

a 1 =

a2 (

cm)

Page 276: Seismic Assessment of Reinforced Concrete Frame Structures

252 Chapter 6

The layout of reinforcement distribution for the beams is shown in Figure 6.3, where the

amount of top and bottom steel are indicated along rows, respectively, above and below each

beam. The adopted depths for the gravity centres of the top and bottom layers (a2 and a1) are

also provided. The amount of slab reinforcement is not included in that figure as it can be

obtained from Figure 6.2. The transversal reinforcement is the same for all storeys and is indi-

cated below the frames in Figure 6.3.

Figure 6.4 includes the reinforcement for columns, where the total longitudinal steel is indi-

cated on the right side of each column and the stirrup distribution is included in the left side.

For each beam and column the steel at the end sections and at the mid-span section are given.

However, for modelling purposes, only the end section data was used since, in the flexibility

element, uniform characteristics are assumed along each left and right portion.

6.2.2.2 Concrete

The concrete characteristics have been considered in accordance with the compressive

strengths listed in Table 5.1. The values of mean cylinder compressive strength there

included were obtained by , where the conversion factor was estimated by inter-

polation between the values for the concrete classes closer to the experimental strength

. For a given class was taken as the ratio of the nominal cylinder to cube strengths

(e.g., for the class C25/30, the conversion factor was ).

The tensile strength , also listed in Table 5.1, was obtained by the EC2 expression

, where the characteristic strength (in MPa) is estimated by .

The so obtained value refers to the axial tensile strength, which was then converted to the

flexural tensile strength by the approximate factor , where h stands

for the cross-section height in the bending direction (REBAP (1984)). For the structure under

analysis the factor takes the value of 1.1.

In the absence of further testing characterization, the concrete elasticity modulus was also esti-

mated by the EC2 expression , for in MPa and in GPa. Such

value corresponds to the secant modulus at 40% of the peak compressive stress.

For the stress-strain behaviour of concrete, the diagrams shown in 4.2.3.2 were adopted.

fcm( )

fcm ξfcm cub,=

ξ

fcm cub, ξ

ξ 25 30⁄ 0.83= =

fctm( )

fctm 0.3fck2 3⁄= fck fcm 8–=

fctm

α 0.6 0.4 h4( )⁄+ 1≥=

Ecm 9.5 fck 8+( )1 3⁄= fck Ecm

Page 277: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 253

According to the notation of those diagrams, the peak compressive stress for unconfined

concrete was taken with the values listed in Table 5.1 and for the corresponding strain, ,

a uniform value of 0.002 was adopted. The values were adopted for the limit of tensile

stress in the linear stress-strain diagram, together with the elastic modulus . For the

confined concrete, the confinement factor and the slope of the descending branch were

obtained from the cross-section and transverse reinforcement data. Residual stresses were

taken zero and 20% of the peak stress, respectively, for unconfined and confined concrete.

6.2.2.3 Steel

The steel data for definition of the stress-strain diagrams included in 4.2.3.2, is based on the

characteristics listed in Table 5.2. The values for yielding and ultimate stresses were assumed

constant for all the beams and equal to the average values corresponding to the bar diameters

actually used. Identical procedure was adopted for the columns and the resulting values are

listed in Table 6.2.

Since no further data was available, the elasticity modulus was assumed as 200 GPa and the

uniform strain at maximum tensile force was taken as 10%, a typical value for the

B500S Tempcore steel (Pipa (1993)), leading to the strain hardening also given in Table

6.2.

6.2.3 Skeleton curves for the section model

The trilinear skeleton curves required for the global section modelling were obtained by the

procedure detailed in 4.2.3, using the section and material data described in the previous sec-

tions.

Table 6.2 Mean tensile properties of steel

Longitudinal Transversal

(mm) (MPa) (GPa) (MPa) (%) (GPa) (mm) (MPa)

Beams12

586 200 678 10 0.956

56914 8

Columns14

577 200 669 10 0.95 810 55916

20

fc0

fcm εc0

fctm

fct( ) Ecm

k Zm

φ fsy Es fsm εsm Esh φ fsyt

εsm( )

Esh( )

M ϕ,( )

Page 278: Seismic Assessment of Reinforced Concrete Frame Structures

254 Chapter 6

The axial forces in columns were estimated from the static vertical loads assumed to act simul-

taneously with the seismic forces, whilst in the beams no axial forces were considered. There-

fore, a preliminary elastic analysis was performed, considering only these vertical loads, and

the resulting axial forces were assumed constant throughout all loading stages.

For beams, the trilinear curves fit quite well the results of the monotonic analysis using a fibre

discretization of the section, as shown in the examples included in 4.3 and in Arêde and Pinto

(1996).

For columns, such fitting depends on the axial force level and on the presence of internal layers

of steel between the main reinforcement layers. Specifically, the last can lead to an underesti-

mation of the yielding capacity, which can unduly advance the formation of plastic hinges in

the columns. This can be seen in Figure 6.5, where, for a lateral column section of the first sto-

rey in the external frame, the trilinear moment-curvature diagram is shown and compared with

the one obtained from fibre analysis.

Figure 6.5 Moment-curvature diagram for a column section. Comparison of trilinear curve and

fibre analysis results.

In order to have a better definition in the yielding zone, all column sections were modelled by

fibre analysis and the corresponding moment-curvature diagrams were obtained and plotted

together with the trilinear one. By visual inspection, better estimates of the yielding moments

CURVATURE (m-1) .00 .04 .08 .12 .16 .20 .24 .28 .32 .36 .40

.00

.40

.80

1.20

1.60

2.00

2.40

2.80

3.20

3.60

4.00x1.E2

Trilinear

MOMENT (kN.m)

Fibre

Original

TrilinearAdjusted

Page 279: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 255

were obtained from the fibre analysis diagram, and corrected trilinear curves were adopted

with new yielding and ultimate points, but preserving the original cracked and post-yielding

stiffnesses; an example of the corrected curve is also depicted in Figure 6.5. The overall

increase of yielding moments due to this correction was about 12% of the original values.

The average ratio of post-yielding stiffness to the cracked stiffness is 0.65% for the beams and

0.7% for the columns.

6.2.4 Hysteretic behaviour

For the hysteretic behaviour modelling, the parameters controlling the unloading stiffness deg-

radation , the pinching effect and the strength degradation , had to be defined. The

following paragraphs explain the related options and assumptions.

6.2.4.1 Unloading stiffness degradation

By means of the shear-drift diagrams, the experimental results show that unloading stiffness

degradation occurs and, therefore it shall be taken into account. However, the adoption of

values is quite difficult since there are no available expressions, relating with the section and

the member characteristics. The authors of the model (Kunnath et al. (1990)) provided only a

range of values, between 2 and 4, adequate for well detailed sections, i.e. in cases where no

high degradation is expected. In other circumstances, it is suggested that the identification of

be based on experimental results.

In the present case, such direct identification is not possible, since no internal forces were

measured and, therefore, no local force-deformation diagrams are available at the element

level. Furthermore, the parameter in the context of the model adaptation as referred in 3.5, is

different from the original one, the difference being dependent on several factors such as the

cracked and the post-yielding stiffnesses, the maximum deformation reached and the amount

of degradation, i.e., the itself. Typically, the common point (as shown in Appendix B) has to

be placed farther away from the origin in the modified model (see 3.5.1) in order to achieve the

same unloading stiffness as in the original model; this means that higher values shall be

adopted.

Although the shear-drift diagrams reflect other phenomena than the stiffness degradation, they

α( ) γ( ) β( )

α

α

α

α

α

α

Page 280: Seismic Assessment of Reinforced Concrete Frame Structures

256 Chapter 6

still provide a means of approximately checking to which extent this issue is affecting the

response. Therefore, upon several trials with different estimates of , while keeping the other

hysteretic parameters fixed, the best agreement with experimental shear-drift and deformation

energy diagrams was obtained by adopting in the first floor beams and columns and

in the remaining members. The low value was found adequate to indirectly

account for the curved part of unloading branches near the zero force level, which reflects the

low stiffness when cracks are open. In all the subsequent calculations, namely for the infilled

frame tests and for the final cyclic tests, the same values of were adopted.

6.2.4.2 Pinching effect

The pronounced pinching exhibited in the shear-drift diagrams is thought to be the result of the

reinforcement asymmetry existing in the beam sections, which became overlapped by rebar

slippage inside the joint concrete core after the peak drifts.

The modelling of pinching due to reinforcement asymmetry was done by adopting the values

as proposed in Appendix B, i.e., given by the ratio of yielding moments of the two bending

directions, which led to values around 0.55.

By contrast, the adequate inclusion of rebar slippage contribution to the pinching effect is a

much more difficult task. It requires, either a specific modelling of the joint behaviour, such as

the one proposed by Monti et al. (1993) for example, or a point hinge modelling at each ele-

ment end section with a moment-rotation law strictly associated with the increment of rotation

due to pull-out. None of these techniques were adopted in the present study, but they are recog-

nized as important to adequately capture the observed pinching effect.

6.2.4.3 Strength degradation

Strength degradation was taken into account by means of the parameter given by the empir-

ical expression included in Appendix B. The confinement and the tension reinforcement ratios

were directly obtained from the cross-section characteristics, whilst the normalized axial force

was estimated from the static vertical loads referred to in 6.2.3. The average values for this

parameter were 2.5% for the columns and 3% and 2%, respectively for positive and negative

bending direction of beams.

α

α 1=

α 4= α 1=

α

γ

β

Page 281: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 257

6.2.5 Damping

In the present work, where damping forces were to be included in the dynamic equilibrium

equation Eq. (5.1), the viscous damping matrix was considered given by the well known

Rayleigh expression (Clough and Penzien (1975)) as follows

(6.1)

where and are constants of proportionality to the mass and stiffness matrices.

This damping matrix is orthogonal to the vibration modes because the mass and stiffness

matrices also verify the orthogonality condition. The parameters and can be obtained so

as to satisfy prescribed damping values for two distinct modes, usually the first and the second.

Let these modes be characterized by the frequencies and , to which the damping factors

and are imposed. The parameters and are given by the solution of

(6.2)

and, therefore, the corresponding matrix satisfies the desired damping in the modes with

frequencies and . For modes with frequencies different from and , the damping factor

cannot be user-controlled since it becomes fixed when the parameters and are set.

The adequacy of using a viscous type damping is as a controversial issue, because if the behav-

iour models were able to “exactly” describe the hysteretic dissipation of energy, there would be

no need for considering other sources of dissipation. Thus, the use of viscous damping forces

to account for energy dissipation appears as an approximation introduced to compensate for

lack of model accuracy.

For the behaviour model used in the present work, two distinct stages can be considered as far

as dissipation is concerned, namely before and after the yielding of critical sections. Before

yielding, the model is typically low-dissipative because it is origin-oriented for unloading and

reloading phases (thus with no dissipation at all) and the energy dissipated due to the cracking

transition is rather low. It follows that viscous damping forces might be needed to simulate the

energy dissipation in dynamic calculations, which was confirmed by the comparison with

C

C amM akK+=

am ak

am ak

f1 f2

ξ1 ξ2 am ak

4πf1( ) 1– πf1

4πf2( ) 1– πf2

am

ak⎩ ⎭⎨ ⎬⎧ ⎫ ξ1

ξ2⎩ ⎭⎨ ⎬⎧ ⎫

=

C

f1 f2 f1 f2

am ak

Page 282: Seismic Assessment of Reinforced Concrete Frame Structures

258 Chapter 6

experimental results included in 6.4.4. Therefore, for the low level test, the numerical analysis

was performed using the experimentally obtained damping factor of 1.8% for both the first and

the second modes.

For loading levels after yielding, the model is dissipative, but several questions can be asked:

• How accurate is the dissipation capacity of the model and, if insufficient, how much of the

energy should be dissipated by viscous forces?

• If viscous damping is to be used, which factor(s) should be adopted, taking into account the

progressive damage and the modifications of the dissipation capacity of the structure? Actu-

ally, it does not seem rational to adopt the same damping factor as before yielding, because

it was referred to an almost undamaged structure and proved to be suitable for the low-dissi-

pative model stage.

These aspects are further discussed in 6.4.4, but it can be anticipated here that good results

were obtained by considering no viscous damping forces for the numerical simulations of the

high intensity tests.

The stiffness matrix and the corresponding frequencies were adequately updated for the initial

conditions of each test in order to have a damping matrix more “coherent” with the actual state

of the structure, i.e. externally unloaded but with the residual deformations and the damage

induced by the previous testing phase.

6.2.6 Modelling of infills

Infill panels were modelled based on a detailed work by Combescure (1996), where multi-

level analyses are used for the simulation of the non-linear response of infilled frames. Basi-

cally this work consisted on the following steps:

• Development and/or improvement of a set of numerical tools for the non-linear analysis of

masonry infill panels, both at local (fine) and at global levels.

• Application of these tools to the numerical analysis of a single infilled frame experimentally

tested at LNEC (Pires F. (1990)), aiming at: a) a better understanding of the behaviour and

strength mechanisms, as well as the influence of modelling strategies and related parame-

ters, and b) checking the adequacy of global modelling and the procedure for parameter

identification based on results of the local modelling.

• Application to the full-scale infilled frames under study in the present work, in order to: a)

Page 283: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 259

assess the influence of boundary conditions on the infill behaviour by means of local mod-

elling of panels both in an isolated frame and in the complete structure, and b) compare the

results of the global modelling against the experimental ones.

6.2.6.1 Setup of numerical tools for infill modelling

For the local (fine) analysis each constitutive material was modelled by a specific behaviour

law, some of them already available in CASTEM 2000, and briefly described in the following

paragraphs. Further details can be found in Combescure (1996).

The model proposed by Menegotto and Pinto (1973) was used for simulating the behaviour of

steel reinforcement, geometrically supported by uniaxial bar elements.

The concrete of the frame was simulated in plane-stress conditions; a non-linear cracking

model (the so-called Ottosen model, Dahlblom and Ottosen (1991)) was considered for the

tensile behaviour, while a linear elastic model was adopted for compression, because the main

source of concrete non-linearity is considered to arise from the cracking phenomena. The

cracking model is of “fixed-crack” type with the maximum principal stress criterion for first

crack development; it accounts for the tensile softening of concrete and the anisotropy induced

by crack formation; additionally, it incorporates a regularization method (Hillerborg et al.

(1976)) to reduce the dependency of global results on the mesh-size.

The interface between the concrete frame and the infills is of great importance for the tested

structure behaviour because significant contribution for non-linearity arises from that zone.

Two types of modelling strategies were available and considered for this purpose, namely the

unilateral contact problem solution and joint elements.

• The unilateral contact between two solids is modelled in CASTEM 2000 by means of kine-

matic conditions, handled by the Lagrange multiplier method, and leads to neither tensile

nor shear force transfer if contact is deactivated.

• The available joint elements (Pegon and Pinto (1996)) can be used in 2D analysis, where the

generalized stresses are the normal and shear forces across the joint plane and the general-

ized deformations are the corresponding relative displacements. The behaviour can be sim-

ulated either by an elastic-perfectly-plastic dilatant model (both for normal and shear

forces) or by a non-dilatant elasto-plastic model with softening for tensile and shear forces.

The latter allows to consider tensile strength progressively decreasing to zero as the joint is

Page 284: Seismic Assessment of Reinforced Concrete Frame Structures

260 Chapter 6

opening, and thus avoids an overestimation of the panel resistance if the tensile strength is

assumed constant.

Masonry infill panels under cyclic loading, are subjected to fast and strong degradation of their

mechanical characteristics, namely the compressive strength. Obviously, also the tensile

strength degrades rapidly; however, since cracking mainly occurs at the frame-panel interfaces

where the non-linear behaviour is modelled by the joint elements, the panel behaviour can be

assumed mainly controlled by compressive stresses. Therefore, a 2D plasticity-based model

was developed (Combescure (1996)) allowing cyclic loading and strength degradation to be

taken into account in the principal stress space; two plasticity surfaces are considered, defined

according to the lowest or the highest principal stress. An internal surface accounts for the

elastic domain (with kinematic hardening), while an external one defines the maximum

strength of the material (with isotropic negative hardening); the latter is affected by the plastic

energy dissipated due to the internal surface, which simulates the effect of cyclic loading in the

strength degradation.

The global analysis consisted of the two following approaches:

• Each infill panel was modelled by a pair of diagonal struts ruled by a uniaxial force-defor-

mation behaviour law, while each member of the reinforced concrete frame was modelled

by one internal linear elastic element and two finite length plastic hinge elements with a

moment-curvature relationship as given in Appendix B.

• Each storey (frame plus infill panels) was modelled by a single shear-beam element ruled

by a shear-drift law based on a primary trilinear curve (with softening branch) and a set of

hysteretic rules.

In the present work context only the infill modelling by diagonal struts is relevant, in order to

be used together with the flexibility elements for the frame members. No tensile resistance is

assumed for diagonal struts, while for compressive behaviour the strut model is based on an

axial force-deformation multi-linear law schematically shown in Figure 6.6 where both the

model basic curve and an arbitrary cyclic path (labelled with numbers next to the arrows) are

included; the model main features are as follows:

• For monotonic loading, the initial elastic branch (OC) is followed by the cracking phase

(CP), during which cracks in the frame-panel interface take place. Then, a plastic zone (PS)

is adopted to simulate the masonry crushing at the compressed corners; it holds until a cer-

Page 285: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 261

tain maximum deformation is reached and then a softening branch (SM) is enforced.

• For hysteretic behaviour, unloading or reloading is origin-oriented in the cracking phase

(thus the non-linearity is only due to stiffness decrease); once in the plastic or softening

branches, unloading is done at constant stiffness (the full-cracked one, OP) until null resist-

ance is achieved; this is kept during the sliding phase until reloading starts, whose rules can

account for the typical pinched shape of infill diagrams due to a pronounced delay of crack

closure. The strength degradation due to cyclic loading is also reproduced by affecting the

maximum restoring force previously reached by a reduction factor which depends on the

accumulated plastic deformation.

Figure 6.6 Diagonal strut model

Both the primary curve and the model parameters for hysteretic behaviour are obtained from

results of refined analyses with the local models, which provide information about the horizon-

tal shear force in the panel (T), the corresponding distortion (drift) and average vertical

panel deformation . By considering the distortion and vertical extension as the most

significant panel deformation modes, the diagonal axial force (N) can be related with T and the

corresponding axial deformation can be written in terms of and , by means of

geometrical conditions. This compatibilization of generalized forces and deformations

between the two modelling approaches allowed to define a procedure (Combescure (1996)) for

the identification of global modelling parameters based on refined modelling technique.

6.2.6.2 Application to a single frame

Experimental tests carried out at LNEC on single infilled frames were simulated by Combes-

cure (1996), using the refined modelling technique above referred, and the following conclu-

sions have been extracted:

• From all the available interface modelling tools, the joint model with softening leads to bet-

εaxial

N

C

P

M

S

816

7

3

1110

12 14

5

1

2

13

4

15

6

17 18

9

19

Sliding phase

StrengthDegradation

O

γ( )

εvertical( )

εaxial( ) γ εvertical

Page 286: Seismic Assessment of Reinforced Concrete Frame Structures

262 Chapter 6

ter estimates of initial stiffness and allows to simulate the stiffness decrease due to cracking

at the interfaces prior to masonry crushing in the panel corners;

• Compression diagonal struts clearly form between opposite panel corners and the maximum

strength is reached when masonry crushes in the diagonal extremities.

• In the near-failure phase, the diagonal tends to move and secondary diagonals are engaged,

due to stress transfer from the crushed corners to the less damaged neighbourhood; thus,

one has to bear in mind the fixed-diagonal model limitations for simulating highly damaged

panels.

The use of diagonal modelling, with the primary curve fitted to the results from refined analy-

sis under monotonic loading, showed shear-drift curves of cyclic tests with the same aspect of

those from the refined analysis; additionally, if strength degradation is accounted for, reasona-

ble agreement with experiments can be obtained. The procedure for obtaining the primary

curve could be tested but the identification of hysteretic parameters still lacks further experi-

mental support.

6.2.6.3 Application to the full-scale structure

The four-storey full scale structure tested at ELSA was modelled using the refined analysis

technique (Combescure (1996)) subjected to an inverted triangular force distribution on the

uniformly infilled configuration.

Additionally, several sub-structures were also modelled and subjected to horizontal displace-

ment-controlled loading; these sub-structures consisted of one or two panels with the surround-

ing frame members, respectively, for the isolated panel configuration and for the total storey

configuration, and for both the first and the second storeys.

The main aim was to compare the behaviour of a given panel in an isolated configuration with

that in the whole structure. Deformed shapes, plastic deformation distributions and shear-drift

diagrams of infill panels were compared in both situations, allowing to draw the following

main conclusions (Combescure (1996)):

• Panels considered in the structure are subjected to vertical tensile or compressive stresses,

induced by the over-turning moment, that cannot be captured by an isolated panel analysis;

hence, the compression diagonal assumption is not strictly valid. However, shear-drift dia-

grams are not very different from those in the isolated or storey configurations and they still

Page 287: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 263

can be used to define the overall panel behaviour.

• Beams on the panel lower boundary play an important role on the panel behaviour and,

therefore, must be taken into account for a better assessment of the panel characteristics.

The seismic behaviour of the whole structure was then simulated by means of global model-

ling. Reinforced concrete members were modelled, as in previous studies (Pinto et al. (1994)),

by the association of two plastic hinge elements and a linear elastic one. The infills were mod-

elled by diagonal struts, the primary curve having been obtained from the local refined analysis

of the first and second storey panel sub-structures. The remaining upper storeys were consid-

ered with the same characteristics of the second one. Hysteretic behaviour and strength degra-

dation parameters were taken with the values which gave the best simulations for the LNEC

specimens. Note that diagonal characteristics were obtained by means of numerical modelling

rather than experimental results, which means that possible defects inherent to the refined

modelling technique will obviously affect the global modelling.

Results have shown a good description of the experimental base shear but with a clear under-

estimation of the top displacement. This appears to be due to stiffness over-evaluation in the

less damaged storeys (the third and fourth ones) where the analytical shear-drift diagrams are

fairly poor approximations to the experimental response.

6.2.6.4 Infill panels in the present study

The infill modelling is not an item of major concern in the present work context and is

included here mainly for the sake of completeness. Thus, the same global modelling of infills

by diagonal struts was adopted here, exactly as used in Combescure (1996) where details can

be found. However, from the results there included, it is recognized that further effort should

be put in the definition of panel characteristics, particularly those related to the initial stiffness

and interface modelling in the upper storeys. It is worth noting that the adopted diagonal char-

acteristics were based on the local (refined) analysis and, according to Combescure (1996),

they led to less good results than a cruder modelling of global storeys by shear-beams but with

parameters directly extracted from the experimental results.

A more accurate definition of diagonal characteristics is beyond the scope of this work, but this

modelling deficiency has to be kept in mind when analysing the comparison of numerical sim-

ulations with experimental results included in 6.4.

Page 288: Seismic Assessment of Reinforced Concrete Frame Structures

264 Chapter 6

6.3 Damage quantification

6.3.1 General overview

The structural damage induced by earthquake action must be quantitatively defined, in order to

assess how close is the state of structure in relation to a pre-defined set of limit states.

The structural damage is usually seen as the ratio of a demand quantity (obtained from the

structural response to seismic events) to the ultimate capacity (Park et al. (1984)), leading to a

damage index. The demand is expressed in terms of one or more response variables, the so

called damage parameters (e.g. displacement, curvature, energy, etc.).

It is noteworthy that damage parameters in this context differ from the concept of damage var-

iables used in Continuum Damage Mechanics, in which nonlinear behaviour is explicitly

dependent on the damage state of the structure and vice-versa. Typically, in such formulation,

the damage variable works as an “internal variable” explicitly taken into account in the consti-

tutive model, whilst in the present context, the damage parameters for damage index definition

refer to “output variables”.

Damage indices can be defined at several levels, namely at the structure, the element, the sec-

tion or even the fibre level. The level for damage definition cannot be finer than the discretiza-

tion level, in the sense that, for example, if the non-linear behaviour of a building structure is

associated to one DOF per floor, the damage index cannot be computed in terms of member

deformations or plastic hinge rotations.

The damaging process in reinforced concrete elements under earthquake loading is usually a

combination of large strain excursions with several repetitions of load reversals. Damage

parameters related to the first phenomenon are often the peak values of response variables; typ-

ical examples are the storey displacements or the plastic hinge rotations. On the other hand, the

strength and stiffness degradation induced by load reversals, particularly those of high ampli-

tude, cause failure for deformation demands lower than those under monotonic loading. There-

fore, adequate damage indices for cyclic conditions shall be based in damage parameters able

to “memorize” the history of deformation, which can be achieved by means of cumulative

measures of the response variables. The total dissipated energy or the cumulative inelastic

deformation (or ductility) are examples of such damage parameters.

Page 289: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 265

Thus, damage indices can be defined in terms of peak values, or cumulative measures or even

combination of both. Extensive overviews of damage indices published in the literature are

presented by Coelho (1992) and particularly by Fardis et al. (1993) who discuss several pro-

posals for damage indices which are categorized, analysed and applied to a large set of monot-

onic and cyclic tests, under uniaxial or biaxial loading, most of them up to failure. The

comparison of test results against damage predictions using those indices allowed their assess-

ment and even led to a new proposal of an energy-based damage index. Such a detailed analy-

sis of the available indices is deemed unnecessary in the present work; nevertheless, brief

references are included next, essentially based in Fardis et al. (1993), concerning the most

widely used indices.

The typical form of damage indices based on response peak values (usually of displacements

or deformations) is the following

(6.3)

where is the damage parameter, is the available capacity and is the threshold of ,

above which the damage is supposed to start. The exponent accounts for the rate of increase

of damage index with the damage parameter and the threshold is often considered as the

yielding displacement. Obviously, if the peak value for is entered in Eq. (6.3) the maxi-

mum damage index value is obtained.

For monotonic loading such index seems adequate since Eq. (6.3) yields for

and when reaches the available capacity. By contrast, no measure of cyclic effects is

included, which renders it inadequate for assessing the damage state when cyclic deterioration

is present.

Another way of expressing a peak value based damage index is by means of the ratio of secant

flexibilities as proposed by Roufaiel and Meyer (1987), where , and

are, respectively, the secant flexibilities at the current deformation, at yielding and at ulti-

mate capacity in monotonic loading.

Attempts to define a damage index in terms of cumulative inelastic deformation have been first

based on the low and high cycle fatigue of materials and are given by

Dδ δth–δu δth–-----------------⎝ ⎠

⎛ ⎞m

=

δ δu δth δ

m

δth

δM δ

D 0= δ δth=

D 1= δ

D f fy–( ) fu fy–( )⁄= f fy

fu

Page 290: Seismic Assessment of Reinforced Concrete Frame Structures

266 Chapter 6

(6.4)

where is the amplitude obtained in the ith half-cycle of the response (supposing the

response previously divided into equivalent half-cycles) and b is a constant. However,

experimental evidence supports the use of Eq. (6.4) for steel but not for reinforced concrete.

Damage indices based on cumulative energy parameters can also be used and are often defined

using the dissipated energy normalized by the potential energy up to failure in monotonic load-

ing . An example of such proposal was introduced by Meyer et al. (1988), in which two

indices are calculated separately for each loading direction and then combined into a unique

one; this proposal has the particular feature of also including the energy dissipated in interior

cycles as a normalizing quantity (i.e. summed to ) and thus reducing the influence on the

damage index of the occurrence of a great number of interior cycles, as demonstrated in Coe-

lho (1992).

Experimental evidence suggests that a damage index should combine contributions from both

peak and cumulative responses. The first proposal of such a damage index was introduced by

Banon (1980) and refined by Banon and Veneziano (1982), through a non-linear combination

of , the total dissipated energy normalized by the deformation energy up to yielding, with

the so-called Flexural Damage Ratio (FDR), given by where refers to the

secant flexural flexibility at peak deformation. The study was supported by a small set of

cyclic tests, to which the proposed index was applied leading to normally distributed results

with mean 1.0 and coefficient of variation 27.5%.

Along the same trend line, Park et al. (1984) presented a mixed damage index, nowadays

widely used and known as the Park and Ang index, given by

(6.5)

or by

(6.6)

Dδi

δu-----⎝ ⎠

⎛ ⎞b

i 1=

N

∑=

δi

N

Eu

Eu

En

FDR fm fy⁄= fm

D δM

δu------ β

Qyδu----------- Ed∫+=

D δM

δu------ β δ

δu-----⎝ ⎠

⎛ ⎞ α EdEc δ( )-------------∫+=

Page 291: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 267

where and have the meaning above introduced, is the yield strength, is the

incremental hysteretic energy, is the hysteretic energy per cycle at deformation , and

and are non-negative parameters. All the parameters in Eqs. (6.5) and (6.6), namely , ,

, and were empirically obtained by statistical regression of a large set of experimen-

tal tests (142 monotonic and 261 cyclic) performed up to failure. Eqs. (6.5) and (6.6), along

with the empirical expressions for the parameters, were applied to the performed test set and

the authors found log-normal distributions of with identical characteristics in both cases,

i.e., mean value 1.0 and coefficient of variation of 54% for the first and 50% for the second.

Therefore, due to its simplicity, Eq. (6.5) appears preferable for the seismic damage assess-

ment; details about the determination of the corresponding parameters, as proposed in (Park et

al. (1984,1987a), Park et al. (1987b)), are presented later in this work.

Fardis et al. (1993) proposed a new energy-based damage index, with a similar structure to that

of Park and Ang index, and given by

(6.7)

where stands for the maximum deformation energy over the response to earthquake

loading, is the total deformation energy up to failure under monotonic loading and the

remaining terms have the same meaning as for the Park and Ang index. Note the closeness of

these two indices: the term is a measure of the maximum deformation , since this

deformation is involved in the computation of that energy; similarly is essentially

dependent on the deformation at monotonic failure.

Damage assessment by Eq. (6.7) has the advantage of handling multiaxial deformations, since

energy is a scalar which can be computed in terms of force and displacement or deformation

vectors. This is of particular interest for the earthquake response in three dimensions, particu-

larly in columns due to biaxial bending plus axial force interaction. However, the so-obtained

damage index refers to the whole element and not to individual plastic hinges at element end

sections, which may become a serious drawback if mechanism-based reliability analysis is

sought (where damage must be evaluated at the plastic hinge level). To overcome this short-

coming, the authors suggest energy splitting between both element end sections, which still

involves some problems as pointed out in Fardis et al. (1993) and discussed later in 6.3.2.4.

δM δu Qy dE

Ec δ( ) δ α

β δu Qy

Ec δ( ) α β

D

DmaxEd β Ed∫+

Ed u,------------------------------------=

maxEd

Ed u,

maxEd δM

Ed u,

Page 292: Seismic Assessment of Reinforced Concrete Frame Structures

268 Chapter 6

Concerning the parameters involved in Eq. (6.7), namely the factor and the failure deforma-

tion under monotonic loading for computation of , the expressions for the Park and

Ang index parameters were used and their adequacy was discussed and assessed by the appli-

cation of Eq. (6.7) to the large experimental data set above mentioned (which included some

biaxial bending tests). Results of Eq. (6.7) showed quite good agreement with experimental

ones, mostly when a constant value of 0.03 was adopted instead of the expressions proposed

by Park et al. The best choice of parameter fitting led to a mean damage value of 1.05 with a

coefficient of variation of 55%, which is still a very high result scatter, demanding further

improvements on parameter assessment.

The damage index proposal given by Eq. (6.7) seems to be a very promising and consistent

option for damage quantification, the major problem being the link with subsequent reliability

studies. The output provides an overall picture of the damage distribution in the structure, and

even a global average damage value, but the lack of information about the damage state of the

critical zones (plastic hinges) appears as a weak point of the proposal which needs to be

improved.

From an assessment of damage indices performed by Fardis et al. (1993), it became clear that

the best available are the Park and Ang proposal and their own suggestion of energy-based

index described in the same study. In the present work, the Park and Ang damage index given

by Eq. (6.5) was preferred, for the following reasons:

• In spite of the promising results of the proposal by Fardis et al. (1993) and of its agreement

with experimental results (as good as the Park and Ang expression agreement), Eq. (6.5) is

still the most established and widely accepted at present in damage assessment.

• Since only planar analyses are performed, with no bending and axial force interaction, the

damage source is restricted to uniaxial bending and, therefore, the potential advantage of

using Eq. (6.7) is diminished.

• It is of low computational cost in the context of the developed computer code and, more

importantly, the damage values are directly associated with each plastic hinge as they come

from end section deformations;

In the following paragraphs details are given of the expressions adopted for the chosen peak

value measure and the specifically related parameters.

β

δu Ed u,

β

Page 293: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 269

6.3.2 The Park and Ang damage index

6.3.2.1 The damage parameter

In the present study, Eq. (6.5) was used to quantify damage in terms of the chord rotation at

each element end section. Being defined as the rotation between the tangent to the element axis

at one end section and the chord connecting both end sections, it follows that chord rotations

coincide with the rotations and in the element reduced space (or basic system) as shown

in Figure 2.4-e). Particularly, since only planar analyses are performed, the rotations of interest

are , i.e. those producing bending in the frame plane; in the following, they are simply noted

by . Therefore, the damage index as used in the present work is expressed by

(6.8)

where, the yielding moment is known from the model skeleton curve and the parameter ,

the same as used in the hysteretic model for controlling the strength degradation, is given in

Appendix B by Eq. (B.4). The peak value is directly obtained from the maximum response,

for each element end section, whilst the remaining terms, namely the ultimate rotation and

the dissipated hysteretic energy , are computed as explained in 6.3.2.3 and 6.3.2.4.

Damage values are calculated for both positive and negative bending directions, thus requiring

, , and to be evaluated separately. The hysteretic energy contributes equally to the

positive and negative bending damage indices, for consistency with the definition of parameter

as stated in Appendix B. Indeed, according to Figure B.3, the incremental hysteretic energy

refers to a complete cycle (possibly with interior small ones) instead of any energy splitting

due to positive and negative deformations.

Rotations were preferred to curvatures because they can give an integrated measure of damage

associated with each end section and, most importantly, they are less sensitive to the plastic

hinge development length and to the distribution of curvatures there existing. If, by virtue of

the assumed flexibility distributions, the curvature presents a very sharp distribution in the end

section neighbourhood, then very high curvature values may be found and the damage

becomes overestimated. Moreover, in reinforced concrete elements and due to the crack devel-

opment, the curvature concept has to be understood in an “average” manner, i.e., as the rotation

θy θz

θy

θ

D θM

θu------ β

Myθu------------ Ed∫+=

My β

θM

θu

Ed∫

θM θu My β

β

Page 294: Seismic Assessment of Reinforced Concrete Frame Structures

270 Chapter 6

divided by the finite distance between two sections where deformations are obtained. There-

fore, the rotation damage appears closer to the physical phenomenon than curvature damage

and can be better compared with available experimental results, most often expressed in terms

of rotations or displacements.

It is worth recalling that the option for chord rotation as damage parameter is inspired on the

equivalence of an element in anti-symmetric bending with the two cantilevers having the same

chord rotations of each end section, as shown in Figure 6.7.

Figure 6.7 Equivalence of element in anti-symmetric bending with cantilever elements

Each cantilever length is given by the so-called shear-span , where stands for

the end section shear force. The tip displacement equals the displacement of the point of

inflection I (for ) relative to the tangent to the element axis at the end section and the

chord rotation becomes the cantilever drift ; however, one must bear in mind that

such equivalence is strictly valid only for anti-symmetric bending, because in the other cases

the inflection point is not necessarily lying on the undeformed axis.

Note that the chord rotations are directly known in the flexibility element context; in addition,

although the above stated analogy is not general, it is helpful to overcome the difficulty of

defining yielding and ultimate chord rotations, indeed the major drawback of adopting chord

rotation as damage parameter.

1 2

M1

M2+

-

+M2

-M1

I

Ls1

Ls2

L

Ls1

Ls2θ1

θ2

θ2

θ1δ1

δ2

δ1

δ2

1 2

LsiMi Vi⁄= Vi

δi

M 0=

θi δi Lsi⁄

Page 295: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 271

6.3.2.2 Yielding rotation

The need for defining the yielding rotation is mainly due to ductility assessment purposes.

Beside the curvature ductility factor given by , where is the yielding curvature,

the chord rotation ductility factor (or simply rotation ductility) can be defined by .

Similarly to , also the yielding rotation can be associated with the limit state of tensile

reinforcement yielding at the most stressed section, in general the end sections. However,

while the attainment of that limit state is sufficient for the yielding curvature definition, the

quantification of the yielding rotation requires further knowledge about the deformation along

a certain length, which, by the analogy shown in Figure 6.7, can be assumed as the shear-span.

At this stage, two important aspects must be discussed: which shear-span shall be considered

and in what consists the deformation along that span?

The shear-span is obviously dependent on the moments installed at both end sections when

yielding is attained at the end section of interest. From the point of view of the stand-alone ele-

ment, this is very unpredictable, since it depends on factors that are unknown prior to the struc-

ture analysis, such as the initial distribution of internal forces and subsequent redistributions

due to plastic hinging, the configuration of applied loads, etc. Moreover, the presence of trans-

verse loads applied along the element complicates even more the problem, because the validity

of “splitting” the member behaviour into the two cantilevers becomes questionable. Taking

into account all the involved uncertainties causing fluctuation of the shear-span value, and

bearing in mind the essentially conventional character of the shear-span, the half member

length is considered for the shear-span of all element end sections, regardless of the bending

direction and of the existence, or not, of transversal loads.

While recognizing that such option can be questionable, as it may not be fully consistent with

the element state when the maximum value of chord rotation is reached during the response, it

still provides a means of comparing, for all the structure elements, how far the chord rotations

go with respect to a conventional measure of yielding.

Regarding the deformations along the shear-span, they include several contributions, namely

the flexural and shear, elastic and inelastic, deformations, the effects of cracks, inclined or not,

the bond-slippage effects, the tension-stiffening effect between cracks, etc. In the context of

the Park and Ang damage model (Park et al. (1984)), the authors considered the yielding rota-

µϕ ϕ ϕy⁄= ϕy

µθ θ θy⁄=

ϕy θy

Page 296: Seismic Assessment of Reinforced Concrete Frame Structures

272 Chapter 6

tion given by

(6.9)

where the subscripts “flex”, “slip” and “shear” indicate the source of each contribution, i.e.,

flexural, slippage and shear, respectively, and the superscript “PA” stands for Park and Ang.

The flexural contribution results from the assumption of a linear variation of curvature along

the shear-span (from at the end section to zero at the inflection point) and also includes the

elastic contribution of shear.

The term is estimated as a concentrated rotation due to reinforcement slip at the face of

the crack often occurring at the end section. A bond-slip model is adopted, similar to the one

proposed in the CEB-FIP Model Code (1990), to obtain the slip when the tension reinforce-

ment yields; the corresponding rotation is calculated by , where is an estimate

of the internal lever arm.

The contribution of is roughly estimated by an idealized shear cracking model, where

45o inclined cracks are assumed to develop, uniformly spaced by , along a certain length

from the end section to the inflection point. At each crack a constant concentrated rotation

is assumed to occur, leading to a transverse displacement of the inflection point, equal to the

rotation times the mean distance to the crack. An approximate expression is proposed for the

length of the “no shear crack zone” and thus, by summing up the contributions due to all the

cracks, the transverse displacement of the inflection point can be expressed in terms of ,

as well as the chord rotation .

Empirical expressions were obtained for by means of the large set of experimental results

used for the damage index validation (Park et al. (1984)), specifically the tests where the yield-

ing point could be identified and the experimental chord rotation could be measured.

After substitution of in the left hand side of Eq. (6.9) and upon calculation of the terms

and , the shear contribution was obtained and, subsequently, was esti-

mated allowing regression analysis.

The full set of expressions for the application of Eq. (6.9) is extensively described in Park et al.

(1984) and was later detailed and translated to more usual European notation and units by

θyPA θy flex, θy slip, θy shear,+ +=

ϕy

θy slip,

sy

θy slip, sy z⁄= z

θy shear,

z

θs

δs θs

θy shear,

θs

θyexp

θyexp

θy flex, θy slip, θy shear, θs

Page 297: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 273

Fardis et al. (1993). Note that the peak response displacement as used by Park et al. (1984)

was calculated by means of force-displacement relationships tuned by the experimental results,

namely the trilinear envelopes and the hysteretic energy absorption per cycle. This led to struc-

tural responses in which all the above mentioned non-linear phenomena are included since the

model was fitted for that purpose. Therefore, both the displacement (or rotation) demand and

their yielding and ultimate values are consistently obtained.

In the present work the adopted model does not take into account the slippage and inelastic

shear contributions for the member deformations. Thus, it appears logical that for the evalua-

tion of the chord rotation ductility the yielding rotation shall include only the flexural (elastic

and inelastic) and the elastic shear contributions as for the calculation of the response values of

. According to the trilinear model used, with the adjustments for the cracking plateau as

referred in 3.5, and for the assumption of inflection point at mid-span as stated above, the

yielding chord rotation is obtained by integration of the shaded diagram of curvatures shown in

Figure 6.8 over an element with half of the total length; the steps referred in 3.6.7 are applied

for due consideration of elastic shear deformation.

Figure 6.8 Yielding chord rotation

6.3.2.3 Ultimate rotation

The ultimate chord rotation necessary for the damage index evaluation by Eq. (6.5) was

first proposed by the authors (Park et al. (1984)), as given by

(6.10)

where the ultimate rotation ductility was given by empirical expressions obtained from

regression analysis of experimental results, and calculated using Eq. (6.9). Another empir-

ical expression was proposed later (Park et al. (1987a)) for a one step calculation of ,

δM

θ

L/2 L/2

Moments Curvatures

Myϕy

ϕcc

ϕc

CY CY

Mcθy

Yielding

θu

θuPA µθ u,

PA θyPA

=

µθ u,PA

θyPA

θuPA

Page 298: Seismic Assessment of Reinforced Concrete Frame Structures

274 Chapter 6

directly from the member characteristics and is given by

(6.11)

where is the volumetric confinement ratio, is the normalized axial force (positive for

compressive forces), is the mechanical ratio of tension reinforcement and is the cylindri-

cal compressive strength of concrete. and refer to the shear-span and to the effective sec-

tion depth, respectively.

From the study performed by Fardis et al. (1993), it was concluded that the best agreement of

the Park and Ang index with the experimental results was achieved by using Eq. (6.11) for

definition.

However, the direct use of Eq. (6.11) for calculation of in this work does not seem very log-

ical for the same reasons pointed out before. The response value of provided by the model

does not include the whole set of non-linear phenomena implicitly taken into account in the

experimentally based expression of and, therefore, underestimated damage values are

likely to be obtained if such expression is directly used.

Actually, an alternative procedure for consistency between the analysis and the damage model

assumptions could be based on Fardis et al. (1993), in which the skeleton curve of the point

hinge model, directly expressed in terms of chord rotations, was adjusted to match and

, in the yielding and the ultimate points, respectively. However, such fitting of the skeleton

curve is not possible with the model used herein, because it is based in curvatures and no pre-

defined plastic hinge length is considered to provide “equivalence” between curvature and

rotation.

Another solution to estimate could be the integration of the curvature diagram as proposed

for the yielding rotation. The ultimate limit state, characterized by the ultimate point in the tri-

linear curve, is assumed to be attained at the end section and a linear moment diagram is

adopted between that section and the inflection point as shown in Figure 6.9; the correspond-

ing curvature diagram can be integrated to obtain the chord rotation at that end section.

However, such chord rotation is a sort of lower bound of the ultimate rotation for similar rea-

θuPA 0.0634

Ls

d-----⎝ ⎠

⎛ ⎞0.93 max ρw 0.004( , )

max ν 0.05( , )-----------------------------------

0.48ωt

-0.27fc-0.15=

ρw ν

ωt fc

Ls d

θu

θu

θ

θuPA

θyPA

θuPA

θu

Page 299: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 275

sons as for the yielding case; indeed, the non-linear effects of reinforcement bond-slippage and

pull-out and diagonal cracking caused by inelastic shear are not included and they are even

more pronounced near failure. Moreover, the ultimate rotation in the context of seismic resist-

ant structures often corresponds to a certain residual strength still retained after the peak

strength deformation is exceeded. Such is exactly the assumption underlying the proposals of

Park et al. (1984) for quantification of , in which failure is conventionally defined at

0.80Mmax on the descending branch of the monotonic moment-rotation curve.

Figure 6.9 Estimation of ultimate chord rotation by curvature integration

For the structure under analysis, the ultimate rotations were first calculated by integration of

curvatures, assuming the inflection point at the mid-span section, and the obtained damage

indices were excessively high, having reached maximum values about 0.75 for the simulation

of the 1.5S7 test. According to the damage index authors (Park et al. (1987a)) such value cor-

responds to a severe degree of damage, beyond repair, with extensive spalling and crushing of

concrete and buckling of rebars. The physical appearance of the structure after that test did not

reveal such extent of damage; indeed it appeared in a fairly good shape.

For these reasons such estimate of ultimate rotations could not be accepted and, considering

that direct application of Eq. (6.11) is not suitable as above explained, an intermediate

approach was adopted as follows.

Both the ultimate and yielding rotations were first calculated, according to Park et al., by Eqs.

(6.9) and (6.11) allowing a ultimate ductility factor to be obtained as . This

factor was preferred, instead of the one included in Eq. (6.10) for which direct expressions

were also proposed in Park et al. (1984), for the above explained reasons of better adequacy of

Eq. (6.11). Then, the ultimate rotation is obtained by

θu

Ultimate

Moments Curvatures

ϕy

ϕu

θu

My

Mc

Mu

CYU CYU

L/2 L/2

µθ u, θuPA θy

PA⁄=

Page 300: Seismic Assessment of Reinforced Concrete Frame Structures

276 Chapter 6

(6.12)

meaning that the same ultimate ductility is kept from the empirical expressions, but applied to

the yielding rotation obtained by the same procedure as used for the response evaluation with

the adopted model. Regarding the shear-span in Eq. (6.11), the same assumption was adopted

as for the calculation, i.e., half of the total element length for each end section.

While recognizing that such damage quantification procedure is debatable due to its somewhat

heuristic nature, it must be recognized that this is still an open issue whose solution is far from

straightforward. Indeed, a more coherent approach would require the explicit quantification of

bar-slippage and inelastic shear effects, but that is beyond the present work scope. Therefore,

rather than an absolute and accurate evaluation of damage at a given critical zone, the presently

adopted damage quantification should be mostly regarded as a comparative measure between

damaged zones (or between different structures); nevertheless, the results of its application to

the four storey structure under the 1.5S7 test (as presented later in 6.4) led to much more rea-

sonable damage values (maximum values around 0.4) and, thus, have encouraged its adoption

for damage quantification purposes in the present work.

6.3.2.4 Hysteretic dissipated energy

For the computation of the energy dissipated by hysteresis, the procedure adopted by Fardis et

al. (1993) was used in this study. The section internal forces and deformations relevant for hys-

teresis dissipation consist on moments and curvatures , whose deformation energy,

upon integration along the element length , leads to the total deformation energy of the ele-

ment.

However, considering the internal moments at each element node (say and ), in equilib-

rium with the internal section moments , the following expression applies

(6.13)

The splitting of element energy into the two end sections is convenient for damage computa-

tion purposes, but it is not a straightforward task. Indeed, it is a problem of similar nature to

that of defining the adequate shear-span for computing yielding and ultimate chord rotations,

θu µθ u, θy flex, θuPA θy flex, θy

PA⁄( )= =

θy

M x( ) ϕ x( )

L

M1 M2

M x( )

Ed∫ M x( ) ϕ x( )d xd0

L

∫∫ M1 θ1d M2 θ2d+( )∫= =

Page 301: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 277

since that splitting could be done by dividing the integration length into two sub-zones

between each end section and the inflection point, the result being assigned to each element

end zone. However, despite the fact that such integration is a minor problem in the present

flexibility formulation context, this way of splitting fails whenever the inflection point falls

outside the element, or when there is more than one such point (as is the case when transversal

loads are considered). Specific criteria are required to split the energy integrated between the

two end sections for the first case, or the amount of energy in between inflection points for the

second case.

Alternatively, the energy splitting can be done as given by the right-hand side integral of Eq.

(6.13), by assigning each term inside brackets to the corresponding end section, thus

avoiding the computation of the integrals along the member length. This procedure is inde-

pendent of whether the inflection point exists or not and, in general, the results agree with the

distribution of deformations. However, this is not problem-free as explained next:

• It can be shown that the result of is not coincident with the energy integration

between the end section i and the inflection point (if existing), except for perfectly anti-

symmetrical situations; nevertheless, this means that, as long as the behaviour of the two

end sections is not very different, the deformation energy can be split by this simpler proce-

dure.

• In some particular situations, unreasonable results can be obtained by , as is the case

of Figure 6.10, where the illustrated beam is bent by the action of moment , the moment

diagram being shown in Figure 6.10-b) and corresponding to the incremental curvature dia-

gram depicted in Figure 6.10-c); by integration of along the ele-

ment, one would assign a non-zero energy value to the right node given by

, whilst a null value is obtained for and, consequently, an overesti-

mation of energy for the left node given by . Note that a similar inconsistency would

occur for a flexible support at node 2, allowing positive rotation but still for negative

moment ; a negative energy would be obtained by against the evidence of posi-

tive energy as the result of integration.

In spite of these shortcomings, this way of splitting energy was preferred to the integration pro-

cedure, for the following reasons:

• The main aim is to estimate the dissipated energy by hysteresis, whose contribution to the

damage value is weighted by the factor , having low values, typically less than 0.1; there-

Mi θid( )

Mi θid

Mi θid

M1

dE x( ) M x( ) ϕ x( )d xd=

dE2 dE x( )l1

L∫= M2 θ2d

M1 θ1d

θ2d

M2 M2 θ2d

β

Page 302: Seismic Assessment of Reinforced Concrete Frame Structures

278 Chapter 6

fore, any errors in estimating the energy contribution do not significantly affect the result;

note that the same reason would not apply if the fully energy-based index given by Eq. (6.7)

would had been adopted.

• In view of the previous point, the non-dependency on the existence of inflection point

makes this procedure advantageous, specially if one bears in mind that the energy is well

computed; indeed, if errors exist, they affect only the splitting between the end sections, but

the total energy is assigned to the element, meaning that any underestimation of damage at

one end section is balanced by overestimation at the other end.

Figure 6.10 Example of inconsistency in energy splitting between element end sections:

a) Beam and deformed shape, b) bending moments and c) curvature diagrams.

In a complex structure, where non-linear effects may lead to redistribution of internal forces

during earthquake loading, the type of particular situations as shown in Figure 6.10 is very

unpredictable. Therefore, since it is not acceptable that a given plastic hinge ends-up with a

negative value for dissipated energy, if such is found at the end of the analysis, then the corre-

sponding energy is set to zero.

M1

dϕ x( )L

I

1

a)

2

l1

M x( )

M2

dθ1dθ2 0=

x

I

dϕ1

dϕ2

M1

M2

b)

c)

+

+

Page 303: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 279

6.4 Analysis of results from numerical simulations

6.4.1 General

The results of numerical simulations with the flexibility global element model are analysed in

the following sections through systematic comparison with experimental results. The main

objective is the assessment of the numerical model ability for reproducing the behaviour of

complete structures, taking advantage of pseudo-dynamic full-scale tests as an excellent means

of analysis of the mechanical response.

Additionally, the flexibility modelling performance is also compared to other analytical model-

ling tools. Among these, the analysis by means of point hinge or fixed length plastic hinge

models is, for sure, the most widely used at present. The latter modelling strategy is more ade-

quate in view of inelasticity spreading and, therefore, it was adopted in this comparative study;

for simplicity it is briefly denoted by F.H., standing for Fixed Hinge length.

Due to the nature of the PSD testing technique, all the experimental phases (except the free-

vibration tests) consisted of quasi-statically imposing displacements at each storey of the struc-

ture, which means that, from the strict experimental point of view, the tests are essentially tra-

ditional quasi-static ones since no actual inertia or damping effects developed in the structure.

Thus, for a given testing phase, numerical static analyses can be performed for an external

loading consisting of the displacements really imposed to structure, and the obtained response

can be compared to the experimental one without involving additional dynamic effects.

For adequate analytical simulations of the dynamic behaviour, three major issues have to be

properly addressed, viz the hysteretic model, the vibration frequencies and the damping. These

issues are fairly inter-related, since a good assessment of frequencies depends on the adopted

model, and the need for explicit inclusion of viscous damping depends on the model ability to

simulate the dissipation characteristics.

Naturally, dynamic modelling is more demanding and difficult than static analysis, because,

even if a suitable model of the hysteretic behaviour is available to provide good estimates of

peak response values, a wrong assessment of frequencies or an inadequate characterization of

damping (if not included in the behaviour model) may lead to poor simulations of the structural

dynamic response.

Page 304: Seismic Assessment of Reinforced Concrete Frame Structures

280 Chapter 6

In view of the above issues, several analyses were performed depending on their nature (static

or dynamic), on the type of test and on the modelling option. These analyses are summarized

in Table 6.3 where the actually performed numerical simulations are indicated with a cross.

A major concern of the numerical simulations was to follow the real test sequence as closely as

possible. Therefore, the actual conditions of load application and, particularly, the unloading

phases between tests had to be adequately traced out as described in 6.4.2.

Results obtained with flexibility modelling are first discussed for static analysis in 6.4.3, in

order to address the model performance for simulating the global hysteretic behaviour under

quasi-static loading. Then the corresponding dynamic analysis results are presented in 6.4.4,

restricted to those indicated in Table 6.3.

The comparison of results from the flexibility and the F.H. modelling is presented and dis-

cussed in 6.4.5, for both static and dynamic conditions.

6.4.2 Procedure for static analytical simulation of the tests

For the static analyses, the boundary conditions had to be adapted for imposing displacements

at each storey level. Therefore, horizontal supports were considered at each floor level, as

shown in Figure 6.11-a), and time histories of the experimentally applied displacements are

prescribed at these supports. The corresponding reactions, recorded at each step, provided the

storey restoring forces from which the inter-storey shears could be obtained and compared to

the experimental ones.

At the end of the test, actuators were driven to zero force, possibly with non-zero residual dis-

placements. In order to analytically simulate this force controlled unloading phase, the hori-

Table 6.3 Numerical simulations performed

Structure Status Test

Flexibility Modelling F.H. Modelling

Static Dynamic Static Dynamic

Bare0.4*S7 x x x x

1.5*S7 x x x x

InfilledUniform x - - -

Soft-Storey x - - -

Bare Final Cyclic x - - -

Page 305: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 281

zontal supports had to be removed from the analytical model, while keeping the structural

stiffness unchanged. Such task is not easily handled in most of the structural analysis computer

codes, in which boundary conditions prescribed at the beginning of the analysis cannot be

modified because the stiffness matrix becomes definitely affected by the stiffness of supports.

However, due to the modularity and object-oriented features of CASTEM2000 (CEA (1990)),

various sources of stiffness can be considered, namely the material stiffness (strictly associated

with the constitutive material) and the rigidity due to local boundary conditions (only those of

fixing supports) and to non-local ones (i.e., those consisting of relations between several

degrees of freedom). The constraints associated with the boundary conditions are handled by

the Lagrange multiplier method (Pegon and Anthoine (1994)), leading to an expanded set of

degrees of freedom consisting on the original kinematic unknowns plus the Lagrange multipli-

ers. The process works independently of the structural material stiffness in the sense that, if

any changes are introduced in the boundary conditions only the coefficients of the Lagrange

multipliers are modified and the material stiffness matrix is not affected; both stiffness contri-

butions are put together in the global matrix only when the system of equations is to be solved.

It is worth mentioning that the possibility of prescribing non-local boundary conditions was

used to simulate the assumption of rigid floor diaphragm, by setting an equality condition for

horizontal displacements (in the loading direction) of all the nodes existing on the same floor,

thereby avoiding the use of artificial rigid links often adopted to connect frames in this kind of

planar analysis. Therefore, the internal and external frames are always shown disconnected, as

their connection is intrinsically taken into account in the model.

Taking advantage of the above referred features, the unloading phase is simply performed as

schematically illustrated in Figure 6.11-b), and consists in:

• reading the last reactions at the floor horizontal supports;

• removal of these floor supports and application of the previous horizontal reactions at each

floor level, progressively decreasing to zero force at completion of the unloading phase.

Before a new testing phase starts, the horizontal supports have to be set up again in order to

apply the new horizontal displacement time histories. Note that the material stiffness is the one

corresponding to the externally unloaded structure, although with a possible internal stress

state due to plastic deformations. Thus, the “actual” state of the structure is kept between con-

Page 306: Seismic Assessment of Reinforced Concrete Frame Structures

282 Chapter 6

secutive testing phases, so that the whole experimental campaign can be simulated as closely

as possible.

Figure 6.11 Storey displacement prescription and unloading to zero actuator forces

The infilled frame tests required special procedures for the simulation. After the unloading

phase of the previous test (the 1.5S7 one), the diagonal struts simulating the infill panels were

added to the unloaded structure with the actual stress, strain and damage states. Again, this

operation was possible thanks to the object-oriented features of CASTEM2000.

From the code point of view, the reinforced concrete frames are represented by one (or more)

model-type object, linked to a geometrical support (the mesh-type object) and a field of mate-

rial properties and characteristics (an object of type field-by-element) in which all the model

internal variables are included. These objects are the strictly necessary to define and construct,

at any stage, the stiffness-type object consisting on the element and the structural stiffness

matrices, either the initial elastic or the tangent ones.

Therefore, other objects of the same type (model, mesh and characteristics) can be defined for

the new sub-structure consisting of the diagonal struts and the corresponding stiffness-type

object can be obtained. These objects are then “assembled” to the corresponding ones of the

bare structure and the analysis is finally performed. Note that the mesh for the infill diagonal

struts is, obviously, built-up with the same nodes of the frames. The procedure is schematically

shown in Figure 6.12 for the uniformly infilled test.

Before the analysis starts, the horizontal supports are again set up exactly as before. Since, in

d1

d2

d3

d4

1

2

3

4

R1

R2

R3

R4

di

(Experimental)Input Output

Ri

t tf t

1

2

3

4

Input Output

R4f

R3f

R2f

R1f

d4r

d3r

d2r

d1r

tu t

Rif

- residual displacements

- tangent material stiffness

and internal forces

a) Imposing storey displacements b) unloading to zero actuator forces

Page 307: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 283

the code context, the boundary conditions lead also to stiffness-type objects, the inclusion of

such supports is simply done by “adding” objects of the same type. That is why, from the user

point of view, the manipulation of boundary conditions turns out to be such an easy task. Once

the adequate stiffness-type object is defined, it can be simply included in or removed from the

total stiffness-type object.

Figure 6.12 Introduction of the infill panel diagonal struts

After the uniformly infilled test was statically simulated by imposing the experimental floor

displacements, again the horizontal supports were removed and the analytical model was

driven to zero external load. However, due to internal stress state, the infill diagonal struts still

interact with the reinforced concrete frames, which means that a straight removal of the infill

stiffness would lead to an out-of-equilibrium state.

Thus, another special procedure is required for the infill removal, as shown in Figure 6.13 and

it is performed as follows:

• the internal forces at the diagonal struts are evaluated from their internal stress state;

• upon removal of the infill contribution for the global stiffness-type object, the external bare

frame is loaded at the nodes where connection with infills did exist before, by forces ;

• the equilibrium being in this way assured, the bare structure is subjected to , progres-

sively decreasing to zero.

Internal Frame External Frame

Internal Frame External Frame

Internal Frame External Frame

Object Type:

Model

Mesh

Field ofCharacteristics

mb

Mb

Cb

Stiffness Kb

Model

Mesh

Field ofCharacteristics

mi

Mi

Ci

Stiffness Ki

Object Type:

Model

Mesh

Field ofCharacteristics

mT = mb+ mi

MT = Mb+ Mi

CT = Cb+ Ci

Stiffness KT = Kb+ Ki

Bare Structure

Infill Struts

Infilled structure

Fi

Fi–

Fi–

Page 308: Seismic Assessment of Reinforced Concrete Frame Structures

284 Chapter 6

Figure 6.13 Unloading of infilled frame configuration: a) removal of actuators and b), removal

of infill panels

At this stage, the configuration of the non-uniformly infilled frame could be set up exactly by

the same procedure as for the uniformly infilled case: first the model, mesh, characteristic and

model-type objects were defined for the diagonal struts (now, only at the three upper floors)

and superimposed with the bare frame ones; then the floor horizontal supports were added

again and the static analysis proceeded with the experimentally imposed displacements. At the

end, the same unloading process was applied (as for the uniformly infilled case) before starting

the final cyclic test simulations.

6.4.3 Static analysis by flexibility modelling versus experimental tests

The static analyses included in this section are strictly concerned with the assessment of the

flexibility modelling ability to simulate the hysteretic behaviour of the structure. However,

besides the simulations of experimental tests, a monotonic pushover analysis was numerically

performed in order to have an estimate of both the maximum base shear and the global yielding

displacement.

Since all the following analysis are statically performed, it is worth mentioning that when the

term “time” is used, it actually refers to the step of the analysis.

6.4.3.1 Pushover analysis

The pushover analysis was carried out by applying progressively increasing static forces at

each floor, with an inverted triangular distribution as indicated in Figure 6.14, up to a maxi-

mum top displacement similar to the maximum reached during the tests (about 0.60 m).

External FrameInternal Frame External Frame

Infill StrutsInfilled structure

After testing and unloadingto zero actuator forces

Internal Frame External Frame

Bare structure

Strut internal forces Fi Unloading of

tu t

- Fiinteraction forces

a) b)

0

0

0

0

d4

Page 309: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 285

Figure 6.14 Pushover analysis: inverted triangular force distribution.

The obtained diagrams of base shear-top displacement and of storey shear-drift are plotted in

Figure 6.15, clearly showing a global behaviour of trilinear type. Thick and thin solid lines cor-

respond to pushover for positive and negative displacements, respectively. Due to frame and

beam section asymmetry and to the presence of vertical loads, positive and negative pushover

analysis do not lead to coincident results; however, differences are not significant.

Figure 6.15 Pushover analysis: base shear-top displacement and storey shear-drift diagrams.

Visual estimates of the yielding point can be extracted from Figure 6.15-a) as indicated by

small circles. Depending on the definition of global yielding, different values can be obtained:

• should global yielding be considered as the first attainment of yielding at the section level,

an estimate of 8 cm seems to be adequate;

• if the yielding point is defined where the force-displacement curve clearly changes slope,

Internal Frame External Frame

F1

F2

F3

F4d4

TOP DISPL.(m)

BASE SHEAR (kN)

.0 .15 .30 .45 .60 .75 .0

.15

.30

.45

.60

.75

.90

1.05

1.20

1.35

1.50 x1.E3

DRIFT (%)

STOREY SHEAR (kN)

.0 1.5 3.0 4.5 6.0 7.5 .0

.15

.30

.45

.60

.75

.90

1.05

1.20

1.35

1.50 x1.E3

Storey 1

Storey 2

Storey 3

Storey 4

0.14 m

0.11 m

0.08 m

a) Base Shear - Top Displacement b) Storey shear - drift

NegativePositive

NegativePositive

Page 310: Seismic Assessment of Reinforced Concrete Frame Structures

286 Chapter 6

then the yielding point approaches 14 cm for the top displacement;

• finally for the intersection of the straight lines fitting the post-cracking and the post-yielding

branches, the yielding displacement becomes an intermediate value around 11 cm.

The comparison of these values with the experimental yielding estimate is difficult to perform,

because no experimental pushover test was performed. As already referred, such an estimate (7

cm) was obtained from the base shear-top displacement diagrams of the 0.4S7 and 1.5S7 tests,

therefore with the inherent difficulties of establishing an adequate envelope of the force-dis-

placement diagram. However, in spite of the unavoidable crudeness of the experimental esti-

mate, the obtained analytical values show a trend for yielding displacements higher than the

experimental ones.

By contrast, the maximum base shear appears underestimated compared to experimental

results: according to 5.5.6, the maximum base shear reached for the final cyclic tests is higher

than 1400 kN, whilst in the pushover analysis (in both directions) it is around 1300 kN. This

aspect might be partially due to the low values of post-yielding stiffness adopted for the global

section model (see 6.2.3) and also due to strength mechanisms that may be differently acti-

vated for pushover analysis and for the tests.

Attention is drawn to the fact that, as for the experimental results, also the pushover analysis

led to maximum drift at the second storey as evidenced in Figure 6.15-b) by the dashed line

connecting the storey yielding points. This fact corroborates the reasons already given in 5.5.4

for the occurrence of the maximum drift at the second storey.

6.4.3.2 Bare frame seismic tests

The simulations of the bare frame seismic tests (see 5.5.4) are described in the next paragraphs

and compared to the experimental results. When time-histories or diagrams are compared in

the same drawing, dashed lines always refer to the experimental results.

Low level test (0.4S7)

For the 0.4S7 test, time histories of the first and fourth storey shear, the corresponding shear-

drift diagrams and evolutions of absorbed energy are plotted in Figure 6.16. The base shear-top

displacement diagram and the peak value profile of storey shear are shown in Figure 6.17,

Page 311: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 287

including also a plot of the cracked zones of each structural member and the spatial distribution

of the maximum positive and negative chord rotations.

From the time histories of storey shear a very good agreement is found between analytical and

experimental results. There is a less good agreement during the initial phase before the first

peaks occur, which is related with the difficulty of adequately characterizing the initial stiff-

ness of the members. The presence of an initial state of micro-cracking, related with shrinkage

effects or caused by the specimen transportation, and variations on the slab participation may

justify the higher initial stiffness obtained in the analytical simulation and evidenced in Figure

6.16-b).

However, once larger incursions in the cracked phase have occurred, member stiffness became

well estimated and the analytical response matched the experimental one, particularly in the

first storey. In the top storey, even after the peaks, the agreement was not so good, though fairly

satisfactory, which may be related with the lower drift values obtained there (less than 50% of

the lower storeys) and a more reduced extent of cracking, thus keeping the uncertainty about

the initial stiffness.

Shear-drift diagrams show that unloading from the first peak (around 2.3s) is not followed by

the analytical response, since this corresponds to member sections only cracked for one bend-

ing direction, thus having a higher stiffness than the fully cracked one as assumed in the ana-

lytical model.

Nevertheless, after large peaks in both directions, the assumption of origin-oriented unloading

and reloading stiffness seems to be quite reasonable, as no major residual drifts persist after

significant cracking is stabilized. On the other hand, this is not sustained by the absorbed

energy diagrams where significant discrepancy can be found between the analytical and the

experimental results; yet, notice must be taken that the involved energy level for this test is less

than 10% of the energy absorbed in the 1.5S7 test and, therefore, this error in energy evaluation

is expected to significantly reduce in presence of the dissipated energy at the end of a post-

yielding analysis.

Page 312: Seismic Assessment of Reinforced Concrete Frame Structures

288 Chapter 6

Figu

re 6

.16

0.4

S7 te

st. S

tatic

ana

lysi

s ve

rsus

exp

erim

enta

l sto

rey

resu

lts

Tim

e (s

)

SHEA

R (k

N)

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

-7.5

-6.0

-4.5

-3.0

-1.5

.0

1.5

3.0

4.5

6.0

7.5

x1

.E2

DRI

FT (x

1.E-

3)

SHEA

R (k

N)

-5.0

-4

.0

-3.0

-2

.0

-1.0

.

0

1.0

2

.0

3.0

4

.0

5.0

-7

.5

-6.0

-4.5

-3.0

-1.5

.0

1.5

3.0

4.5

6.0

7.5

x1

.E2

Tim

e (s

)

ENER

GY

(kJ)

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

10.

0

Tim

e (s

)

ENER

GY

(kJ)

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

10.

0

Stor

ey 4

Stor

ey 1

Tim

e (s

)

SHEA

R (k

N)

DRI

FT (x

1.E-

3)

SHEA

R (k

N)

.0

.

8 1

.6

2.4

3.2

4.

0 4

.8

5.6

6.4

7.

2 8

.0

-3.0

-2.4

-1.8

-1.2

-.6

.0

.6

1.2

1.8

2.4

3.0

x1

.E2

-2.0

-1

.6

-1.2

-.

8 -

.4

.0

.4

.8

1

.2

1.6

2

.0

-3.0

-2.4

-1.8

-1.2

-.6

.0

.6

1.2

1.8

2.4

3.0

x1

.E2

b) S

tore

y sh

ear-

drift

dia

gram

c) S

tore

y en

ergy

a) S

tore

y sh

ear

Anal

ytic

Expe

rim

enta

l

Page 313: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 289

Figu

re 6

.17

0.4S

7 te

st. S

tatic

ana

lysi

s ve

rsus

exp

erim

enta

l res

ults

TOP

DIS

PL (x

1.E-

2 m

).

BASE

SH

EAR

(kN)

-5.0

-4.

0 -3

.0 -

2.0

-1.0

.0

1

.0

2.0

3.0

4.

0 5

.0

-7.5

-6.0

-4.5

-3.0

-1.5

.0

1.5

3.0

4.5

6.0

7.5

x1

.E2

SHEA

R (x

1.E2

kN

)

STO

REY

.0

1

.5

3.0

4

.5

6.0

7

.5

1

2

3

4

Neg

ativ

e =

-3.2

4Po

sitiv

e =

4.0

6

a) B

ase

shea

r - to

p di

spla

cem

ent d

iagr

amb)

Pea

k sh

ear s

tore

y pr

ofile

c) A

naly

tical

cra

ckin

g pa

ttern

d) C

hord

Rot

atio

ns. M

ax.(m

Rad

):

Anal

ytic

Expe

rim

enta

l

Page 314: Seismic Assessment of Reinforced Concrete Frame Structures

290 Chapter 6

It is worth analysing the sources of the discrepancy of dissipated energy (given by the inferior

peaks of diagrams): the energy “jump”, between 2s and 3s time, is associated with the non-

fully cracked stiffness of sections where the cracking moments were exceeded only for one

bending direction, and cannot be followed by the numerical model because if cracking occurs

it is assumed in both directions; after that, the dissipated energy discrepancy increases progres-

sively (although at a decreasing rate) due to residual deformations, that, indeed do exist, and

cannot be considered in the model.

Globally, a quite good agreement is achieved between the base shear - top displacement dia-

grams resulting from numerical and experimental analysis (see Figure 6.17-a)), as well as for

the peak values of storey shear shown in Figure 6.17-b), where differences between calculated

and measured values are less than 5%.

From the final cracking pattern (see Figure 6.17-c)) the following is highlighted:

• The behaviour of short-span beams is mainly controlled by the lateral action, since no cen-

tral cracking develops and the extent of cracking decreases from the lower to the upper

beams, thus following the trend of the storey shear profile.

• In the long-span beams, the influence of the vertical load is quite apparent due to the

obtained central cracking, which is less in the external frames due to their stiffness being

more than twice that of the internal one (whilst the vertical load is exactly the double);

moreover, the cracking in these beams is more uniform in height than for the short-span

beams, thus confirming the greater influence of vertical loads for the longer span beams.

• Cracking is more relevant in columns adjacent to the long span beams, due to the vertical

load effect, but is well distributed along the height.

Peak-values of chord rotations (see Figure 6.17-d) are essentially decreasing along the height,

although for the long span beams and the adjacent columns this trend is not strictly followed,

due to the vertical load effect. Note that, both maximum positive and negative peak-values are

higher than experimental ones (max. 2.51 mRad, according to Negro et al. (1994)), but they are

not directly comparable because the latter refer only to the plastic hinge zone.

However, integration of numerically obtained curvatures inside the length of the plastic zone

considered in the test, leads to maximum values of 2.1 and -3.0 mRad, therefore comparing

much better with the experimental ones. Such analytical values actually refer only to the 0.4S7

Page 315: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 291

test, because the contribution of initial rotations previously installed due to vertical loads was

removed to make results comparable with the experimental ones.

High level test (1.5S7)

The analytical and experimental results for the 1.5S7 test are included in Figures 6.18 to 6.20.

For the first storey, it is apparent the quite good agreement between analytical and experimen-

tal shear responses shown in Figure 6.18-a): the highest experimental shear peak is only 10%

underestimated by the analysis. This can be related with the collaborating slab width, for which

the experimental estimates were somewhat higher than those adopted in the calculations

(therefore developing higher strength), and also with the low post-yielding stiffness in the

moment-curvature relationships. By contrast, during reloading phases, the analytical shear

overestimates the response, due to the pronounced pinched shape of the experimental shear-

drift diagram shown in Figure 6.18-b). The analytical loops become exterior to the experimen-

tal ones, thus leading to a larger amount of dissipated energy as evident from Figure 6.18-c).

Actually, the comparison of shear-drift diagrams suggests that it might be difficult to ade-

quately simulate such a pronounced pinching effect by two straight lines as adopted in the

present behaviour model; nevertheless, energy dissipation agrees reasonably well.

In the top storey, deviations between analytical and experimental shear are more evident (peak

values about 16% underestimated) and the absorbed energy obtained from the calculations is

significantly lower than the measured one; however, the effect in the overall dissipation is

minor due to the reduced contribution of that storey. A lower stiffness is exhibited by the corre-

sponding shear-drift diagram, suggesting that excessive cracking might have been obtained in

the analytical model; several factors may have contributed, namely the slab width and the way

how vertical loads were considered (note that such loads are relatively more important in the

top storey than in the others).

From the above explained, the analytic response is characterized by a slight underestimation of

storey peak shear but, overall, a reasonably good agreement of shear-drift diagrams is found,

although with some difficulties in simulating the pinching effect (see also Figure 6.19-a)) lead-

ing to slight overestimates of the global dissipated energy. However, this fact was somehow

expected, since no rebar slippage and pull-out effects are modelled, and they seem to be

responsible for the significantly pinched shape of the diagrams.

Page 316: Seismic Assessment of Reinforced Concrete Frame Structures

292 Chapter 6

Figu

re 6

.18

1.5

S7 te

st. S

tatic

ana

lysi

s ve

rsus

exp

erim

enta

l sto

rey

resu

lts

.0

.

8

1.6

2.4

3.

2 4

.0

4.8

5.6

6.

4 7

.2

8.0

-1

.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

x1

.E3

-2.5

-2

.0

-1.5

-1

.0

-.5

.

0

.5

1

.0

1.5

2

.0

2.5

-1

.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

x1

.E3

.0

.

8

1.6

2.4

3.

2 4

.0

4.8

5.6

6.

4 7

.2

8.0

.0

.2

.4

.6

.8

1.0

1.2

1.4

1.6

1.8

2.0

x1

.E2

.0

.

8

1.6

2.4

3.

2 4

.0

4.8

5.6

6.

4 7

.2

8.0

.0

.2

.4

.6

.8

1.0

1.2

1.4

1.6

1.8

2.0

x1

.E2

Tim

e (s

)

SHEA

R (k

N)

DRI

FT (%

)

SHEA

R (k

N)

Tim

e (s

)

ENER

GY

(kJ)

Stor

ey 1

Tim

e (s

)

SHEA

R (k

N)

DRI

FT (%

)

SHEA

R (k

N)

Tim

e (s

)

ENER

GY

(kJ)

Stor

ey 4

.0

.

8

1.6

2.4

3.

2 4

.0

4.8

5.6

6.

4 7

.2

8.0

-9.0

-7.2

-5.4

-3.6

-1.8

.0

1.8

3.6

5.4

7.2

9.0

x1

.E2

-1.5

-1

.2

-.9

-.

6

-.3

.

0

.3

.6

.9

1

.2

1.5

-9

.0

-7.2

-5.4

-3.6

-1.8

.0

1.8

3.6

5.4

7.2

9.0

x1

.E2

b) S

tore

y sh

ear-

drift

dia

gram

c) S

tore

y en

ergy

a) S

tore

y sh

ear

Anal

ytic

Expe

rim

enta

l

Page 317: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 293

Figu

re 6

.19

1.5S

7 te

st. S

tatic

ana

lysi

s ve

rsus

exp

erim

enta

l res

ults

TOP

DIS

PL (m

)SH

EAR

(x1.

E3 k

N)

STO

REY

1

2

3

4

a) B

ase

shea

r - to

p di

spla

cem

ent d

iagr

amb)

Pea

k sh

ear s

tore

y pr

ofile

c) A

naly

tical

cra

ckin

g pa

ttern

d) A

naly

tical

yie

ldin

g pa

ttern

BASE

SH

EAR

(kN)

-.2

5 -

.2

-.15

-.1

-.

05

.0

.05

.1

.

15

.2

.25

-1

.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

x1

.E3

.0

.3

.6

.9

1.2

1

.5

Max

. Yie

ldin

g Le

ngth

: 0.2

9 m

Anal

ytic

Expe

rim

enta

l

Page 318: Seismic Assessment of Reinforced Concrete Frame Structures

294 Chapter 6

Figure 6.20 1.5S7 test - Static analysis. Spatial distributions of peak values

a) Chord Rotations. Max.(mRad): Negative = -21.3Positive = 24.0

b) Chord Rotation Ductility. Max.: Negative = 4.0Positive = 4.8

c) Damage. Max.: Negative = 0.36Positive = 0.41

Page 319: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 295

The analytical cracking pattern included in Figure 6.19-c) shows extensive cracking progres-

sion both in beams and columns, having even reached the full length in some beams and indi-

cating the fact that this test went far beyond the cracking phase.

The yielding pattern reveals that plastic hinge formation took place mainly at the beams, where

the maximum yielding progression reached about 30 cm, and also, but less evidently at some

column end sections. Note that yielding lengths are measured in the flexible portion of each

member; for the beams this means yielding development starting at the beam-column face,

whilst for the columns the plotted yielding zones start at the beam-axis. Since the beams are 45

cm deep and the maximum plastic length in columns is visibly less than half the maximum

yielding length, it follows that, all the column yielding zones next to joints, actually do not

develop in the member. Therefore, plastic hinges in vertical elements reduce only to the bottom

sections of base-columns as expected.

The maximum plastic hinge length is 0.67h (where h is the beam depth), thus inside the range

of values often adopted (0.5h to 1.5h), and occurs in the first storey long-span beam due to the

combination of vertical load effects with the high shear at that storey. Following the decreasing

trend of shear in height, the yielding lengths reduce towards the upper storeys.

The spatial distributions of maximum chord rotation, the corresponding ductility and damage

are illustrated in Figure 6.20. Concentric lozenges are plotted for each member end section,

with diagonals proportional to the maximum positive and negative values. In order to identify

where member yielding occurs, only ductility values greater than unity are plotted.

Chord rotations are significantly higher (about 60%) than experimental rotations (see Figure

5.9), as already found for the low level test. Again, the integration of numerically obtained cur-

vatures inside the “experimental” plastic zones (45 cm both in beams and columns) leads to

20.7 mRad and -19.4 mRad, for plastic rotations, referring exclusively to the contribution of

the 1.5S7 test. These results are closer to the experimental ones, but still significantly higher,

which agrees with the overestimation of the yielding displacement given by the pushover anal-

ysis; the low values of post-yielding stiffness used in the section modelling may be responsible

for this rotation discrepancy.

However, the overall distribution of chord rotations agrees with the experimentally measured

rotations, showing decreasing values on beams along the height and higher rotations at the

Page 320: Seismic Assessment of Reinforced Concrete Frame Structures

296 Chapter 6

base-columns in accordance with the yielding pattern. The concrete weakness in the third floor

columns (see concrete properties in Table 5.1) is reflected in locally higher rotations, when

compared to those of the second storey columns where the maximum drift occurs.

Ductility distribution exhibits the same pattern of chord rotations and reaches the maximum

value of 4.8. This indicator of how far the structure is from yielding, serves mainly for compar-

ison between the various loading stages. Its consistency with other measures of ductility, as for

example the top displacement ductility, cannot be easily checked since no direct relation can be

derived for complex structures.

Actually, for the simplest case of a cantilever beam loaded by a vertical force at the tip, the dis-

placement ductility factor is coincident with the chord rotation ductility factor and can

be easily related with curvature ductility , either by the model adopted herein or by other

proposals such as in Paulay and Priestley (1992). However, for complex structures, the relation

between a global displacement and any member deflection or rotation is dependent on several

factors which include: the type of deformations involved (i.e., the significance of shear, joint

and any foundation deformations), the strength proportions between members, the loading pat-

tern, etc. Furthermore, the definition of yielding displacement requires the assumption of the

mechanism likely to develop, which, for subsequent loading stages can change due to succes-

sive formation of plastic hinges. It follows that any explicit relation between global and local

ductility factors appears impossible to be defined, unless parametric studies are performed.

In spite of the absence of any helpful relation between structure and element ductilities, it is

worth recalling that this 1.5S7 test roughly led to the top displacement ductility of 3. The value

obtained for the rotation ductility (4.8) confirms the expected trend in the ductile behaviour of

structures, i.e. the ductility demand of any member of a ductile chain is greater than the global

ductility demand, but no other conclusions can be easily drawn.

The damage pattern appears consistent with the rotation and ductility distributions, illustrating

the dissipative mechanism of the structure, mainly based on the three first storey beams and on

the base columns. Globally the damage is well distributed, although with higher values for the

short-span beams, and the dissipation occurs as expected from the capacity design require-

ments. According to the approximate scale of damage proposed by Park et al. (1984), relating

the damage index values of several earthquake damaged structures with the visible damage

µ∆ µθ

µϕ

Page 321: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 297

observed, the maximum value obtained for this test (around 0.4) seems reasonable, although

somewhat high, when compared with the damage state registered during and after the test:

major cracks permanently open at beam-column interfaces, cracking along members and in

joints for peak deflections and no concrete crushing or reinforcement buckling.

6.4.3.3 Infilled frame tests

Results of infilled frame numerical analyses and their comparisons against experimental

results are shown in Figures 6.21 and 6.22, respectively for the uniformly infilled and the soft-

storey cases. A general overview reveals that, by contrast with the previous results for bare

frame tests, a fairly less good agreement with experimental results is obtained for the infilled

configurations.

For the uniformly infilled case, the first storey shear evolution (see Figure 6.21-a)) matches

reasonably well the experimental response and the peak value is properly estimated. In spite of

a higher stiffness for the negative loading direction, the analytical shear-drift diagram (see Fig-

ure 6.21-b)) gives a good prediction of the experimental behaviour in that storey. After the

large drift peaks have occurred in both directions, i.e. after 4.8s, the numerical and experimen-

tal responses compare very well, which is related to the fact that the panel failed (as actually

observed during the test) and the response became controlled by the frame at that storey.

In the top storey, the analytic results significantly deviate from the experimental ones. Clearly

the analytic initial stiffness is overevaluated and, since the response for this storey takes place

in the low drift range (post-elastic but pre-crushing), the shear force becomes fairly larger than

the experimental one. This is thought to be related with the adopted panel characteristics for

the upper storeys: according to Combescure (1996) they have been taken equal to those of the

second storey, under the assumption that it could be representative of the three upper storeys;

however, this is not confirmed by the results obtained here. In addition, the response in the

cracking range is mainly controlled by the masonry-frame interface behaviour, whose initial

state can hardly be assessed with reasonable accuracy. Indeed, the joint characteristics for

interface modelling were adopted with average values (Combescure (1996)), some of them

estimated from numerical simulations of other experiments independent of the present one.

Thus, it is recognized that further effort should be invested in parameter characterization,

mainly directly from the experimental results rather than from numerical refined simulations.

Page 322: Seismic Assessment of Reinforced Concrete Frame Structures

298 Chapter 6

Figure 6.21 Uniformly infilled test. Static analysis versus experimental results

-5.0 -4.0 -3.0 -2.0 -1.0 .0 1.0 2.0 3.0 4.0 5.0 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

Time (s)

SHEAR (kN)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 -2.5

-2.0

-1.5

-1.0

-.5

.0

.5

1.0

1.5

2.0

2.5 x1.E3 SHEAR (kN)

SHEAR (x1.E3 kN)

STOREY

.0 .5 1.0 1.5 2.0 2.5

1

2

3

4 ENERGY (kJ)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 .0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0 x1.E2

Storey 4

Storey 1

DRIFT (%)

Time (s)

-1.5 -1.2 -.9 -.6 -.3 .0 .3 .6 .9 1.2 1.5 -2.5

-2.0

-1.5

-1.0

-.5

.0

.5

1.0

1.5

2.0

2.5 x1.E3

Time (s)

SHEAR (kN)

DRIFT (x1.E-3)

SHEAR (kN)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

a) Storey shear b) Storey shear-drift diagrams

c) Peak shear storey profile d) Energy diagrams

Analytic

Experimental

Analytic (RC frame only)

Page 323: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 299

Figure 6.22 Soft-storey test. Static analysis versus experimental results

SHEAR (kN)

Time (s)

SHEAR (kN)

DRIFT (x1.E-3)

SHEAR (kN)

STOREY

.0 .5 1.0 1.5 2.0 2.5

1

2

3

4

DRIFT (%)

SHEAR (x1.E3 kN)

a) Storey shear b) Storey shear-drift diagrams

c) Peak shear storey profile

Storey 4

Storey 1

d) Energy diagrams

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

-5.0 -4.0 -3.0 -2.0 -1.0 .0 1.0 2.0 3.0 4.0 5.0 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

Time (s)

SHEAR (kN)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 -2.5

-2.0

-1.5

-1.0

-.5

.0

.5

1.0

1.5

2.0

2.5 x1.E3

-5.0 -4.0 -3.0 -2.0 -1.0 .0 1.0 2.0 3.0 4.0 5.0 -2.5

-2.0

-1.5

-1.0

-.5

.0

.5

1.0

1.5

2.0

2.5 x1.E3

Time (s)

ENERGY (kJ)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 .0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0 x1.E2

Analytic

Experimental

Analytic (RC frame only)

Page 324: Seismic Assessment of Reinforced Concrete Frame Structures

300 Chapter 6

The storey profile of peak shear values confirms how crude is the analytical prediction even if

a good estimate of the base shear was achieved (see Figure 6.21-c)). Note the almost uniform

peak shear obtained in the three upper storeys which reflects the similar characteristics adopted

for the corresponding diagonals.

Figure 6.21-d) shows the experimental and the analytical evolutions of deformation energy; for

the analytical results the contribution of the reinforced concrete alone is also included, from

which can be seen the predominance of the infills in the overall response. The low rate of

energy increase (namely after the larger peaks) is a clear indication of the low dissipation

capacity typical of masonry panels under cyclic loading.

For the soft-storey configuration a reasonably good agreement is found between analytical and

experimental results in the first storey (see Figure 6.22-a)), where the resistance to lateral loads

is mainly controlled by the reinforced concrete frame. For the first incursions into larger drift

ranges, the strength is underestimated, as observed already for the 1.5S7 test, and this trend

remains for the post-peak response in the positive direction, apparently due to excessive degra-

dation of strength. Attention is drawn to the fact that the unloading stiffness degrades more in

the analytical prediction than in the experiment, suggesting that the controlling parameter,

which seemed adequate for the previous bare frame tests, may not be appropriate for such

higher drift range; actually, the same fact is confirmed in the final test results.

The behaviour in the top storey exhibits the same problems as for the uniformly infilled case

(see Figure 6.22-a)): the excessive initial stiffness leads to overestimation of storey shear,

which helped to distort the storey profile of peak shear values. Actually, Figure 6.22-c) shows

a strange decrease in shear at the second storey due to force transfer to the upper storeys, which

remained stiffer due to their lower damage.

The deformation energy diagrams illustrated in Figure 6.22-d) almost duplicated the values of

the uniformly infilled case. This is due to the larger engagement of the reinforced concrete at

the first storey, which gives now the main contribution as clearly shown by the analytical

energy diagrams.

Page 325: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 301

6.4.3.4 Final cyclic test

Due to the extent of damage induced by the soft-storey test, some of the first storey member

end zones had to be repaired as referred in 5.5.6. This intervention cannot be easily simulated

in the analysis, thus introducing an increased degree of uncertainty and difficulty on reproduc-

ing the experimental behaviour for the final test. Additionally, at the end of the soft-storey test

the structure position had to be adjusted in order to reduce the visible residual drifts. Difficul-

ties in monitoring this operation and the resetting of displacement transducers to zero at the

beginning of the final cyclic test, prevented the initial position to be adequately assessed for

the numerical analysis.

The above referred reasons raised the question of whether the final test simulation should be

done after the infilled test analyses, or alternatively, by ignoring them and proceeding directly

from the 1.5S7 test. In order to clarify this topic, the analytical results of the Duct. 3 phase of

the final test were compared with the experiment in two distinct situations, viz by considering

or by neglecting the infilled tests in analytical simulations. The results of such a comparison in

terms of shear-drift diagrams, are shown in Figure 6.23 for both the first and the second storey.

The most relevant aspect is the better agreement with the experimental results when tests with

infills are not considered. By neglecting these tests (see Figure 6.23-a)), for both the first and

the second storeys, the response is quite well captured in the negative side where the maximum

drift was previously exceeded.

However, in the positive side, the first storey diagram exhibits further plastification because

the corresponding peak drift for Duct. 3 has not yet been reached; additionally, the peak shear

force becomes higher than the experimental one because the analytical simulation does “see”

the strength and stiffness degradation occurred during the infilled frame tests.

On the contrary, the effect of considering the tests with infills is a drastic reduction of the first

storey stiffness due to the higher drift values previously reached during the soft-storey test.

Note the comparison of peak values of inter-storey drifts shown in Figure 6.24 for the 1.5S7,

the soft-storey and the final Duct. 3 tests, from which it is apparent that, concerning the rein-

forced concrete structure, only the first storey was seriously affected by the infilled frame tests.

Page 326: Seismic Assessment of Reinforced Concrete Frame Structures

302 Chapter 6

Figure 6.23 Shear-drift diagrams for the Ductility 3 phase of final test. Effects of considering or

neglecting infilled frame tests.

After specimen repairing the experimental stiffness was partially recovered, but this cannot be

reproduced by the analytical simulation; consequently, results deviate quite a lot as evidenced

in Figure 6.23-b) for the first storey. Moreover, such deviation is also responsible for the

results of the second storey: both the initial stiffness and the maximum shear force are overes-

timated, which can be due to force transfer from the first to the second (and less damaged) sto-

rey.

b) Considering infillsa) Direct (neglecting infills)

Storey 1

DRIFT (%)

SHEAR (kN)

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

DRIFT (%)

SHEAR (kN)

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

DRIFT (%)

SHEAR (kN)

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

DRIFT (%)

SHEAR (kN)

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

Storey 1

Storey 2 Storey 2

Analytic

Experimental

Page 327: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 303

Figure 6.24 Inter-storey drift profiles for 1.5S7, soft-storey and final Duct. 3 tests

The aforementioned reasons indicate that analytical simulations of the final cyclic tests are bet-

ter achieved if the infilled frame tests are ignored. Even so, the uncertainty about the initial

position of the structure still persists and this has to be kept in mind when comparing analytical

and experimental results in the next paragraphs.

The first and second storey shear-drift diagrams for the Duct. 5 and 8 stages are shown in Fig-

ures 6.25-a) and 6.25-b), respectively. Together with Figure 6.23-a), the comparison of analyt-

ical and experimental shear-drift diagrams shows that the numerical modelling provides less

good predictions than for the previous tests of the bare structure.

From the above mentioned comparisons, the following aspects are highlighted:

• For Duct. 3, the strength degradation appears to be lower than the experimental one, whilst

the unloading stiffness deterioration seems reasonably modelled (note that it was tuned for

the 1.5S7 test, thus for a similar ductility level as for the present one); by contrast, the

pinching effect is not well modelled as already noted for the previous tests.

• For Duct. 5, in the first storey, further plastification occurs in both directions, but the devel-

oped force is lower than the experimental one; the unloading stiffness degrades more than in

the experimental diagram and the transition between the unloading and the reloading phases

tends to shift away from the null force zone. As for the previous ductility level, the strength

degradation is quite low when compared with the experimental results.

• Still for Duct. 5, but for the second storey, further plastification develops as for the first sto-

1.5S7

DRIFT

STOREY

.0 .5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 (%)

1

2

3

4

Soft-StoreyDuct. 3

Page 328: Seismic Assessment of Reinforced Concrete Frame Structures

304 Chapter 6

rey, but the most relevant aspect is the strength drop occurred in the positive direction for

the second cycle. Actually, it is much larger than the degradation in the following cycles,

which may suggest that it can be related with a mistaken deformed shape of the structure at

the test beginning; the “fatter” loops appear in correspondence with the narrower ones of

the first storey, somehow showing force transfer between the two lower storeys.

Figure 6.25 Final cyclic test, Ductilities 5 and 8: first and second storey shear-drift diagrams

• The Duct. 8 level shows, for the first storey, an increase of the positive shear force without

further plastification: such paradoxical result seems to arise again from force transfer from

b) Ductility 8a) Ductility 5

Storey 1

DRIFT (%)

SHEAR (kN)

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

DRIFT (%)

SHEAR (kN)

-7.5 -6.0 -4.5 -3.0 -1.5 .0 1.5 3.0 4.5 6.0 7.5 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

DRIFT (%)

SHEAR (kN)

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

DRIFT (%)

SHEAR (kN)

-7.5 -6.0 -4.5 -3.0 -1.5 .0 1.5 3.0 4.5 6.0 7.5 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

Storey 1

Storey 2 Storey 2

-4.5 -3.6 -2.7 -1.8 -.9 .0 .9 1.8 2.7 3.6 4.5

-4.5 -3.6 -2.7 -1.8 -.9 .0 .9 1.8 2.7 3.6 4.5

Analytic

Experimental

Page 329: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 305

the second storey, where the peak shear force stays well below the experimental one (a sim-

ilar reason as for the Duct. 5 may justify this discrepancy). The unloading stiffness degrades

significantly more than in the experiment and the resulting diagrams become unreasonably

narrow; additionally, the unloading-reloading transition tends to shift away even more from

the null force zone, but, nevertheless, the peak shear force is well estimated.

The increasing deviation of analytical and experimental unloading stiffness for increasing duc-

tility progressively induces less energy dissipation in the numerical analysis, as shown in the

energy diagrams included in Figure 6.26. Indeed, for Duct. 3, the energy is overestimated

because the model cannot simulate the pull-out phenomena; on the other hand, a less adequate

modelling of the unloading stiffness deterioration is responsible for significant underestima-

tion of the dissipated energy for the Duct. 8 level, in which the pull-out effects become less

important because drifts have larger participation of column deformation (less prone to bar-

slippage than beams).

Figure 6.26 Final cyclic test: total energy diagrams

In order to find out the influence of the latter aspect, other calculations were performed with

increased values of the unloading stiffness degradation parameter only in the first storey

columns (i.e., instead of 1 as referred in 6.2.4, a uniform value of 4 was used for all those col-

umn sections). Figure 6.27 shows the first storey shear-drift and the total energy diagrams for

the Duct. 8 test phase, clearly evidencing that the unloading behaviour is much better simu-

lated by the numerical model (see Figure 6.27-a)); consequently, the energy evolution becomes

ENERGY

0 900 1800 .00

.15

.30

.45

.60

.75

.90

1.05

1.20

1.35

1.50 x1.E3

Ductility 8

Ductility 5

Ductility 3

StepAnalytic

Experimental

α( )

Page 330: Seismic Assessment of Reinforced Concrete Frame Structures

306 Chapter 6

closer to the experimental one (see Figure 6.27-b)). Note, however, that the unloading-reload-

ing transition still occurs out of the zero force zone, thus showing no significant influence of

the parameter on that phenomenon.

Figure 6.27 Final cyclic test Duct. 8: results for modified unloading stiffness degradation

In spite of the deviations of analytical and experimental shear-drift diagrams, the peak shear

response is reasonably well estimated. The corresponding storey profiles are included in Fig-

ure 6.28, showing a fairly good agreement of results for Duct. 3; deviations do appear for Duct.

5, which are partially recovered in storeys 1 and 2 for the Duct. 8 level. Note the remarkable

agreement of base shear for all three ductility stages, significantly improved comparatively to

the 1.5S7 test (see Figure 6.19), which means less than 10% of underestimation as provided by

the analysis.

The strange shape of the Duct. 8 shear-drift diagrams suggests some comments regarding

either a possible mistaken deformed shape at the beginning of the test simulation, or assumed

displacements different from those actually applied. In fact, the first “steps” of diagrams in

Figure 6.25-b) show a sudden drop in the second storey, simultaneously with a “high stiffness”

initial loading branch in the first storey. Such effect may be due to an internal strain distribu-

tion not compatible with externally applied displacements, demanding force adjustments in

order to restore compatibility.

α

b) Total energy diagrama) 1st storey shear-drift diagram

DRIFT (%)

SHEAR (kN)

-7.5 -6.0 -4.5 -3.0 -1.5 .0 1.5 3.0 4.5 6.0 7.5 -1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 x1.E3

Storey 1

ENERGY

0 900 1800 .00

.15

.30

.45

.60

.75

.90

1.05

1.20

1.35

1.50 x1.E3

Step

Analytic

Experimental

Page 331: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 307

Figure 6.28 Final cyclic test: profiles of peak values of storey shear

In other words, if by starting from an initial inter-storey deformed shape (as schematically

shown in Figure 6.29), displacements and are assumed to be applied, the corresponding

drift is as shown in Figure 6.29-b), with the same sign of the initial drift; therefore, a load-

ing or reloading process is induced. On the contrary, if the actually applied displacements are

as shown in Figure 6.29-c), the drift takes the opposite direction to the initial one and unload-

ing occurs; thus, the shear force is likely to reduce, as is apparent from Figure 6.23-b) in the

storey 2.

This fact has been specifically pointed out because the final cyclic test was interrupted at the

end of the Duct. 5 phase, in order to allow for several inspections. The interruption was intro-

duced at zero value of the top displacement, followed by structure unloading to zero forces in

the actuators. The restarting process for the Duct. 8 phase was carried out, first by restoring the

displacements to values prescribed at the interruption step and then by proceeding with the

foreseen top displacement history. During that restoring process, the effectively applied dis-

placements might not have been correctly stored and the displacement series may not exactly

match the real ones.

In fact, a detailed inspection of such series in the stop-restart phase (actually not shown in Fig-

ure 5.12) revealed some inconsistency between the first and the second storey displacements

c) Ductility 8a) Ductility 3 b) Ductility 5

SHEAR (x1.E3 kN)

STOREY

1

2

3

4

.0 .3 .6 .9 1.2 1.5 SHEAR (x1.E3 kN)

STOREY

1

2

3

4

.0 .3 .6 .9 1.2 1.5 SHEAR (x1.E3 kN)

STOREY

1

2

3

4

.0 .3 .6 .9 1.2 1.5

Analytic

Experimental

u1a u2

a

∆ua

Page 332: Seismic Assessment of Reinforced Concrete Frame Structures

308 Chapter 6

recorded in the restarting process, which may be the cause of the unexpected features of the

corresponding analytical shear-drift diagrams. If that had not occurred, both first and second

storey analytical diagrams of Figure 6.25-b) would be expected to shift, downwards and

upwards, respectively, in the positive force direction.

Figure 6.29 Influence of assumed displacements different from the applied ones

Finally, it must be referred that at the end of Duct. 8 stage simulation, the maximum yielding

length reached 1.33h in the first storey beams (for h the beam depth), that is twice the value for

the 1.5S7 test. In turn, the peak values of chord rotation ductility and damage reached 18.6 and

1.57, respectively, both for positive bending and occurring in the first storey short-span beams.

Such a high damage value clearly agrees with the near failure stage and the large inter-storey

drifts reached (over 7%), almost uniformly, in the two first storeys.

6.4.3.5 Summary of static analysis results

The most relevant aspects on the static simulation of the tests are briefly summarized in the

following paragraphs.

The 0.4S7 test analysis has given quite good results in terms of storey shear and shear-drift dia-

grams, mainly for stabilized cracked behaviour. However, due to the origin-oriented features of

the global section model, the dissipated energy is not so well simulated.

Identically, for the 1.5S7 bare frame test, good agreement of results was achieved for the shear

forces (maximum deviation around 10% to 15% from the experiment) and for the dissipated

energy, mainly in the storeys with higher deformations. Difficulties in modelling the pinching

due to anchorage slippage led to a slight overestimation of energy. Concerning more localized

deformations, the plastic zone rotations are somewhat overestimated, thus agreeing with the

u1a u1

r

u2a u2

r∆ua ∆ur∆u0

c) Actually applieda) Initial inter-storey b) Assumed deformed shape displacements displacements

Storey 1

Storey 2

Page 333: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 309

higher yielding top displacement obtained by the analysis; such effects may be due to the low

post-yielding stiffness used for the local section modelling, which may also explain the lower

storey shear values. However, reasonable values are obtained for the damage index, in view of

the observed state of the structure, and the respective distribution corresponds to the expected

dissipation mechanism (strong column - weak beam).

For the infilled frame tests, results diverge from the experimental ones, where and while infill

contribution is important, due to overestimation of the initial stiffness of panels. The uniformly

infilled configuration, after significant cracking of the first storey panel, behaved essentially as

a soft-storey structure and the results for that storey became closer to the experimental ones,

actually as exhibited for the irregular (soft-storey) configuration.

In spite of difficulties in the final test simulation due to the repairing intervention and the resid-

ual deformations after the infilled frame tests, the results in terms of shear are well captured,

although slightly (10%) underestimated (still, the low post-yielding stiffness may be the possi-

ble cause). However, the unloading stiffness degradation was found difficult to simulate with

the present model rule; for a wide range of global ductility, meaning an even wider range of

local (curvature) ductility of which the upper bound refers to a near failure stage, the adopted

rule using a common point leads to excessive degradation of the unloading stiffness.

A general overview of the performed simulations reveals good adequacy of the model to fol-

low the evolution of the global stiffness. With the exception of the infilled frame tests, the glo-

bal average stiffness in experimental shear-drift diagrams is reasonably estimated throughout

the several ductility stages.

6.4.4 Dynamic analysis by flexibility modelling versus experimental tests

The adequacy of the presently adopted model to simulate the global hysteretic behaviour was

already assessed by means of the static analysis, and good performances were obtained, mainly

for the bare frame tests. The model adequacy to trace out the dynamic response is now

addressed in this section.

Before the presentation of the dynamic analysis results, calculated values of frequencies are

compared with those measured for several testing stages, as this is relevant for the subsequent

analysis of results.

Page 334: Seismic Assessment of Reinforced Concrete Frame Structures

310 Chapter 6

The characterization of damping as adopted herein is discussed simultaneously with the pres-

entation of results, restricted to the bare frame tests (the 0.4S7 and 1.5S7 ones) due to the

higher level of uncertainty of infill modelling.

6.4.4.1 Comparison of structural frequencies

The frequencies were experimentally obtained at the following testing phases: before and after

the bare frame tests (respectively, pre-0.4S7 and post-1.5S7), before the uniformly infilled test

(pre-Uniform Infill) and before the soft-storey test (pre-Soft Storey). For comparison purposes

only the fundamental frequencies are presented in Table 6.4, along with the analytical values

obtained by means of the tangent stiffness matrix at the unloaded stage.

The initial frequency (pre-0.4S7) is overestimated by the analysis, which reflects the difficul-

ties in characterizing the initial stiffness as already pointed out for the static analysis of the

0.4S7 test. However, after this test, the experimental frequency is quite well captured by the

model, as confirmed by the pair of values identified by pre-1.5S7. The analytical one corre-

sponds to the initial tangent matrix before the 1.5S7 test, whilst the experimental one was esti-

mated by the last top displacement peaks of the 0.4S7 response, because no measurements

were made between the two tests. Such a good agreement of frequencies is in direct corre-

spondence with the previously obtained agreement of static analysis results after having

exceeded the first peaks (see Figure 6.16-a). The same reasoning applies to the frequencies for

the post-1.5S7 case, which indeed do compare very well.

Table 6.4 Structural frequencies. Comparison of measured values with those calculated by flexibility discretization

Structure Status Case

1st Mode Frequency (Hz)

Measured Calculated

Bare

Pre-0.4*S7 1.78 2.23

Pre-1.5*S7 1.27* 1.29

Post-1.5*S7 0.82 0.84

Infilled

Pre-Uniform Infill 3.34 6.67

Post-Uniform Infill ** 1.00

Pre-Soft Storey 1.67 1.66

Post-Soft Storey ** 0.76

Bare Post-Final Cyclic ** 0.58

(* obtained from displacement history; ** not measured)

Page 335: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 311

For the uniformly infilled configuration a very poor frequency prediction is obtained by the

analysis (twice the experimental value), which is clearly related with the overestimation of the

initial stiffness of infills already detected in the static analysis. At the end of that test no fre-

quency measurements were made, but the analytical value is included to provide an additional

indication of the stiffness decrease along the testing sequence. Note that the obtained value

(1.0 Hz) is higher than that for the post-1.5S7 case, which is due to the presence of infills only

slightly damaged in the two upper storeys although strongly damaged in the two first ones.

On the contrary, the experimentally obtained frequency for the soft-storey case is very well

predicted by the numerical model; this may be related with the first mode deflected shape

which mainly involved the first storey deformation, the remaining ones having moved as an

“almost rigid box” supported by the ground floor columns whose stiffness was well captured.

This is corroborated by the frequency agreement at the post-1.5S7 stage and is due to the fact

that the stiffness of the first storey columns has not been significantly affected by the uni-

formly infilled test in which much lower inter-storey drift was reached.

At the end of the soft-storey test, the analytical frequency (0.76 Hz) dropped below the post-

1.5S7 case, mainly as a result of damage in the first storey columns, and after the final cyclic

tests it became less than one third of the initial value.

From the above paragraphs, it can be concluded that, in spite of the difficulties in estimating

the initial frequency, the model is able to adequately follow the frequency evolution of the bare

frame during the loading process.

6.4.4.2 Low-level test on the bare structure

The dynamic analysis for the 0.4S7 tests were performed under the assumptions stated in 6.2.5,

thus for a viscous damping factor of 1.8%, and results are shown and compared with the exper-

imental ones in Figures 6.30 to 6.32.

Time histories of storey displacements are depicted in Figure 6.30-a), for both the first and the

top displacement. Identically, storey shear responses and the corresponding shear-drift dia-

grams are also included in Figures 6.30-b) and 6.30-c), respectively.

Two distinct phases can be observed in these time histories: the first, approximately until 4s, in

Page 336: Seismic Assessment of Reinforced Concrete Frame Structures

312 Chapter 6

which the analytical response was dominated by the overestimation of frequency and could not

reach the first peaks, both in displacement and in force; the second phase, which, after signifi-

cant cracking induced by the negative peak at 4.5s, was characterized by a stabilized frequency

allowing to closely follow the experimental response, although with slight underestimation of

peaks. This is apparent in both the first and the top storeys, and can be also observed in the

shear-drift diagrams; note that experimental drifts could not be reached in the analysis, which

can be due to the higher initial stiffness.

Note, however, that such underestimation of the experimental response appears to be quite

acceptable, since only minor deviations of relevant results are found. Storey profiles of peak

displacements, drifts and shear forces are plotted in Figures 6.31-a), b) and c), respectively,

allowing to conclude that the maximum top displacement is 15% lower than the experimental

one, whilst the peak drift (second storey) and the base shear are underestimated by 5% and

12%, respectively.

By contrast with static calculations, the energy dissipation in dynamic analysis is due to both

viscous damping and hysteretic restoring forces. Moreover, in the static analysis, the external

loading consisted of imposed storey displacements and the deformation energy could be

directly evaluated by the work done by storey reaction forces on the corresponding displace-

ments, since this total external work equals the internal deformation energy.

However, in the dynamic analysis no explicit external forces are considered at storey level (the

external loading is a base accelerogram) and, therefore, the hysteretic deformation energy has

to be evaluated either by the sum of each element energy (directly available from the flexibility

element model), or by the work done by all components of nodal hysteretic forces on the

respective displacements. The energy contribution of viscous damping has to be computed by

the latter process, since it is not defined at the element level; nodal viscous damping forces are

obtained from velocities and the corresponding work done on nodal displacements yields the

contribution to energy dissipation.

Figure 6.31-d) shows the energy responses from experimental and numerical analysis. The

numerical one is plotted both with and without the viscous damping contribution, in order to

assess its weight in the dissipated energy.

Page 337: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 313

Figu

re 6

.30

0.4

S7 te

st. D

ynam

ic a

naly

sis

with

1.8

% v

isco

us d

ampi

ng v

ersu

s ex

perim

enta

l sto

rey

resu

lts

.0

1

.0

2.0

3

.0

4.0

5

.0

6.0

7

.0

8.0

DIS

PL. (

m)

-5.0

-4.0

-3.0

-2.0

-1.0

.0

1.0

2.0

3.0

4.0

5.0

x1

.E-2

SHEA

R (k

N)

SHEA

R (k

N)

Stor

ey 4

Stor

ey 1

SHEA

R (k

N)

-5.0

-4

.0

-3.0

-2

.0

-1.0

.

0

1.0

2

.0

3.0

4

.0

5.0

-7

.5

-6.0

-4.5

-3.0

-1.5

.0

1.5

3.0

4.5

6.0

7.5

x1

.E2

SHEA

R (k

N)

-7.5

-6.0

-4.5

-3.0

-1.5

.0

1.5

3.0

4.5

6.0

7.5

x1

.E2 .0

1

.0

2.0

3

.0

4.0

5

.0

6.0

7

.0

8.0

DIS

PL. (

m)

.0

1

.0

2.0

3

.0

4.0

5

.0

6.0

7

.0

8.0

-5

.0

-4.0

-3.0

-2.0

-1.0

.0

1.0

2.0

3.0

4.0

5.0

x1

.E-2

Tim

e (s

)

Tim

e (s

)

Tim

e (s

)

Tim

e (s

)

DRI

FT (x

1.E-

3)

DRI

FT (x

1.E-

3)

.0

1

.0

2.0

3

.0

4.0

5

.0

6.0

7

.0

8.0

-3

.0

-2.4

-1.8

-1.2

-.6

.0

.6

1.2

1.8

2.4

3.0

x1

.E2

-2.0

-1

.6

-1.2

-.

8

-.4

.

0

.4

.

8

1.2

1

.6

2.0

-3

.0

-2.4

-1.8

-1.2

-.6

.0

.6

1.2

1.8

2.4

3.0

x1

.E2

b) S

tore

y sh

ear

c) S

tore

y sh

ear-

drift

dia

gram

a)

Sto

rey

disp

lace

men

ts

Anal

ytic

Expe

rim

enta

l

Page 338: Seismic Assessment of Reinforced Concrete Frame Structures

314 Chapter 6

Figure 6.31 0.4S7 test. Dynamic analysis with 1.8% viscous damping vs. experimental results

It is apparent that hysteretic dissipation almost stopped after cracking has stabilized, indeed as

already found in the static analysis. On the contrary, by including the viscous damping contri-

bution, the analytical energy dissipation follows the increase rate of the experimental one. This

means that, except for the 2 to 3s time interval, during which an energy jump occurred due to

large cracking progression, the viscous damping factor of 1.8% seems adequate to compensate

the lack of hysteretic damping inherent in the model for stabilized cracking conditions.

In order to find out the influence of viscous forces in the dynamic response, an additional anal-

ysis was performed for zero viscous damping and the corresponding top displacement and

energy time histories are illustrated in Figures 6.32-a) and 6.32-b), respectively.

STOREY

.0 1.0 2.0 3.0 4.0 5.0

1

2

3

4STOREY

.0 1.0 2.0 3.0 4.0 5.0

1

2

3

4

STOREY

.0 1.5 3.0 4.5 6.0 7.5

1

2

3

4

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0

ENERGY (kJ)

.0

2.5

5.0

7.5

10.0

12.5

15.0

17.5

20.0

22.5

25.0

DISPL. (x1.E-2 m)

SHEAR (x1.E2 kN)

DRIFT (x1.E-3)

Time (s)

b) Peak-drift storey profilea) Peak-displacement storey profile

d) Evolutions of total energyc) Peak-shear storey profile

Analytic

Experimental

Analytic (hysteretic only)

Page 339: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 315

Figure 6.32 0.4S7 test. Dynamic analysis with no viscous damping versus experimental results

It can be observed that, up to 2s time, the analytical response remains essentially as before,

confirming the influence of a poor estimate of the initial frequency. However, the first peaks

between 2s and 3s are better captured and, therefore, larger cracking occurs earlier than for the

previous calculations. Thus, frequency becomes closer to the experimental value and the ana-

lytical response is improved in the interval 3s to 4s, after which the lack of damping leads to an

amplified response, although the peak displacement is not very much overestimated. The hys-

teretic energy evolution is identical to that of the analysis with 1.8% damping, although reach-

ing slightly higher values due to the larger displacements involved.

Comparison of Figures 6.30-a) and 6.32-a) allows to conclude that a globally better response is

obtained if 1.8% of viscous damping is included, bearing in mind that the slight underestima-

tion of peaks is related with the initial stiffness. This better agreement seems quite acceptable

and can be explained as follows:

• The experimentally measured viscous damping factor (1.8%) refers to a structural behav-

iour mainly controlled by micro-cracking, which, after accumulation of micro-cracks leads

to the formation of localized (and eventually visible) cracks typical of the post-cracking

stage.

• Therefore, the 1.8% factor can be regarded as representing the viscous equivalent to the hys-

teretic damping inherent in the cracking process, during which the steel (behaving linearly)

b) Evolutions of total energya) Top-displacement

Time (s)

DISPL. (m) .

-5.0

-4.0

-3.0

-2.0

-1.0

.0

1.0

2.0

3.0

4.0

5.0 x1.E-2

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0

Time (s)

ENERGY (kJ)

.0

2.5

5.0

7.5

10.0

12.5

15.0

17.5

20.0

22.5

25.0

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0

Analytic

Experimental

Page 340: Seismic Assessment of Reinforced Concrete Frame Structures

316 Chapter 6

does not contribute to energy dissipation.

• Consequently, better results can be expected if the stiffness drop, due to cracking through-

out the structure, is progressively taken into account (although without hysteretic dissipa-

tion), but in association with viscous damping equivalent to the hysteretic one actually

existing in the real structure.

Other results of dynamic analysis concerning cracking pattern and distributions of chord rota-

tion peak values are not presented since they do not significantly differ from those of static

analysis. However, it is worth mentioning that maximum positive and negative chord rotations

(3.56 mRad and -2.72 mRad, respectively) are somewhat lower than the static analysis ones, in

agreement with the lower drifts and displacements obtained from the dynamic analysis.

6.4.4.3 High level test on the bare structure

The 1.5S7 test dynamic analysis was first performed for the same conditions of the 0.4S7 test,

namely with a viscous damping factor of 1.8%; results are included in Figure 6.33.

Good agreement is found with experimental results concerning the first higher displacement

peaks (up to 3.5s) and the phase of the response (see Figure 6.33-a)). However, after these

peaks the response becomes progressively underestimated by the analysis, for which two pos-

sible reasons can be pointed out:

• Pinching is more relevant after the attainment of large peaks but the analytical model is not

able to follow that effect because it is mainly related with reinforcement slippage inside

beam-column joints. Therefore, the unloading and reloading stiffnesses are higher than the

experimental ones (see Figure 6.33-c)), loops become “fatter” than in the experiment and

the expected displacements become progressively more difficult to reach; note, however,

that the fundamental frequency from the analysis does not seem very different from the

experimental one (as apparent in Figure 6.33-a)) because the global average secant stiff-

nesses are not very different either.

• The viscous damping contribution to energy dissipation might be excessive, as suggested by

the energy curves shown in Figure 6.33-b); actually, the viscous dissipation added to the

hysteretic dissipation (inherent in the model at this post-yielding stage) led to dissipated

energy higher than the experimental one, and the analytical response became overdamped.

Page 341: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 317

Figure 6.33 1.5S7 test. Dynamic analysis with 1.8% viscous damping vs. experimental results

In order to assess the sensitivity of results to the pinching effect and the amount of viscous

damping, additional analyses were performed. Specifically concerning the first issue, the

pinched shape of force-displacement diagrams was artificially made more pronounced by uni-

formly reducing the pinching moments in the local section behaviour laws to 50% of their orig-

inal value, whilst the viscous damping factor was kept the same; results of such analysis are

shown in Figure 6.34.

The top displacement time history (Figure 6.34-a)) indicates that peak values are better esti-

Time (s)

DISPL. (m) .

Time (s)

ENERGY (x1.E2 kJ)

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 -.25

-.20

-.15

-.10

-.05

.0

.05

.10

.15

.20

.25

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 .0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Storey 4

b) Evolutions of total energya) Top displacement

Analytic

Experimental

c) Shear-drift diagram of 1st storey

Analytic (hysteretic only)

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

x1.E3 SHEAR (kN)

DRIFT (%)

Storey 1

Page 342: Seismic Assessment of Reinforced Concrete Frame Structures

318 Chapter 6

mated, though slightly in excess, but the second half part of the response becomes more out-of-

phase. The modification of pinching effect can be detected in the shear-drift diagram, with

cycles narrower than in the previous analysis, and is reflected in the slight decrease of energy

dissipated by hysteresis (Figure 6.34-b).

Figure 6.34 1.5S7 test. Dynamic analysis with 1.8% viscous damping and modified pinching

versus experimental results

Note that this simple way of modifying the pinching effect intended to show the trend of

results, but cannot be seen as a meaningful result itself. Indeed, such modification could be

hardly justified, i.e., there was no specific reason for that 50% reduction and the occurrence of

Storey 1

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 -.25

-.20

-.15

-.10

-.05

.0

.05

.10

.15

.20

.25

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 .0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Time (s)

DISPL. (m) .

Time (s)

ENERGY (x1.E2 kJ)

Storey 4

b) Evolutions of total energya) Top displacement

c) Shear-drift diagram of 1st storey

Analytic

Experimental

Analytic (hysteretic only)

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5 SHEAR (kN)

DRIFT (%)

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

x1.E3

Page 343: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 319

pinching is not uniform throughout the structure because it is dependent on the deformation

level and is fairly related with the reinforcement slippage inside the beam-column joints; fur-

thermore, the latter aspect requires more refined tools (Monti et al. (1993)) for an adequate

modelling.

The influence of viscous damping was assessed by performing other analyses with zero damp-

ing factor, actually as adopted in the experiment; results are illustrated in Figures 6.35 to 6.36.

The top storey displacement time-history, included in Figures 6.35-a), shows a better agree-

ment with the experimental one, than in the previous analysis with viscous damping included

(compare with Figure 6.33-a)). The experimental displacements are not exceeded by the analy-

sis and the response phase deviation becomes less pronounced than that of Figure 6.34-a).

Among the analyses performed, the last two cycles of the present analytical response exhibit

the best approximation to the experimental one, although still far below.

Beside the difficulty of reproducing the pinching behaviour, higher vibration modes seem to

affect the analytical response as evidenced in time histories of the first and top storey shear

depicted in Figures 6.35-b). Mostly during the second half of the response, higher frequency

components are present that counteract the main one; these components could not be damped

out because no viscous damping is prescribed and the model does not seem to account for them

properly. Note that this effect is clearly more pronounced in the top storey shear response,

mainly ranging in the post-cracking stage where the model hysteretic features are less accurate.

Figures 6.36-a), b) and c) show, respectively, the storey profiles for displacement, drift and

shear peak values. Quite good predictions are provided by the analytical model, which just

slightly underestimate the experimental results: the maximum error is about 7% for the peak

base shear, whilst for the peak top displacement it is less than 4%. The evolution of total

energy is included in Figure 6.36-d) where a very good agreement is also found between anal-

ysis and experiment. This result also supports the thought that no viscous damping needs to be

included if the model adequately accounts for the hysteretic behaviour and the intrinsic dissi-

pation of energy.

Finally, as for the 0.4S7 test, the cracking and yielding patterns, the distributions of peak val-

ues of chord rotation, ductility and damage do not substantially differ from those of the static

analysis and, therefore, they are not presented here.

Page 344: Seismic Assessment of Reinforced Concrete Frame Structures

320 Chapter 6

Figu

re 6

.35

1.5

S7 te

st. D

ynam

ic a

naly

sis

with

zer

o vi

scou

s da

mpi

ng v

ersu

s ex

perim

enta

l sto

rey

resu

lts

b) S

tore

y sh

ear

c) S

tore

y sh

ear-

drift

dia

gram

a)

Sto

rey

disp

lace

men

ts

DIS

PL. (

m)

Stor

ey 4

Stor

ey 1

SHEA

R (k

N)

SHEA

R (k

N)

DIS

PL. (

m)

Tim

e (s

) Ti

me

(s)

Tim

e (s

)

DRI

FT (%

)

.0

1

.0

2.0

3

.0

4.0

5

.0

6.0

7

.0

8.0

-

.25

-.2

0

-.1

5

-.1

0

-.0

5

.0

.0

5

.1

0

.1

5

.2

0

.2

5 .0

1

.0

2.0

3

.0

4.0

5

.0

6.0

7

.0

8.0

-

.25

-.2

0

-.1

5

-.1

0

-.0

5

.0

.0

5

.1

0

.1

5

.2

0

.2

5

.0

1

.0

2.0

3

.0

4.0

5

.0

6.0

7

.0

8.0

x1.E

3

-2.5

-2

.0

-1.5

-1

.0

-.5

.

0

.5

1

.0

1.5

2

.0

2.5

x1.E

3

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

SHEA

R (k

N)

SHEA

R (k

N)

Tim

e (s

)

D

RIFT

(%)

-1.5

-1

.2

-.9

-.

6

-.3

.

0

.3

.

6

.9

1

.2

1.5

-9

.0

-7.2

-5.4

-3.6

-1.8

.0

1.8

3.6

5.4

7.2

9.0

x1

.E2

.0

1

.0

2.0

3

.0

4.0

5

.0

6.0

7

.0

8.0

-9

.0

-7.2

-5.4

-3.6

-1.8

.0

1.8

3.6

5.4

7.2

9.0

x1

.E2

Anal

ytic

Expe

rim

enta

l

Page 345: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 321

Figure 6.36 1.5S7 test. Dynamic analysis with no viscous damping versus experimental results

6.4.4.4 Remarks on energy comparison

So far, energy comparisons in the dynamic analysis context have been presented without any

reference made to the energy input. However, different responses obtained from the analysis

and the experiment, mean that energy input to the system is not the same in both cases. Conse-

quently, energy diagrams presented in the previous paragraphs correspond to the energy

absorbed from sources which are not necessarily coincident in the numerical analysis and in

the experiment; this aspect is further addressed in the following paragraphs.

Consider Eqs. (6.14) of the dynamic equilibrium of both the pseudo-dynamic test and the

numerical analysis

.0 .5 1.0 1.5 2.0 2.5

1

2

3

4 STOREY STOREY

STOREY ENERGY (x1.E2 kJ)

DISPL. (m)

SHEAR (x1.E3 kN)

DRIFT (%)

Time (s)

.0 .05 .10 .15 .20 .25

1

2

3

4

.0 .3 .6 .9 1.2 1.5

1

2

3

4

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 .0 .5

1.0 1.5 2.0 2.5 3.0

3.5 4.0 4.5 5.0

b) Peak drift storey profilea) Peak displacement storey profile

d) Evolution of total energyc) Peak shear storey profile

Analytic

Experimental

Page 346: Seismic Assessment of Reinforced Concrete Frame Structures

322 Chapter 6

(6.14)

For simplicity no viscous damping forces are included and the mass matrix is assumed the

same in the experiment and in the analysis. Both displacements , accelerations and restor-

ing forces have the subscript e or a, referring to the experiment and the analysis, respec-

tively.

The corresponding energy balance equations are obtained by integrating Eqs. (6.14) over the

displacement history up to a given instant t, and are given by

(6.15)

The first terms in Eqs. (6.15) can be simplified to and , referring to

the kinetic energy of the system at instant t, in the following denoted by and , respec-

tively. On the other hand, the second terms of the first members refer to the absorbed energy

(including the elastic and the hysteretic components) and are denoted by and . Thus,

Eqs. (6.15) can be re-written as

(6.16)

where and are the input energies, respectively the experimental and analytical ones.

In the energy diagrams shown so far, the experimental curves actually refer to the term and

the analytical ones to . Since input energy terms and are not necessarily identical

because experimental and analytical displacements are not the same, the question can be asked

whether or not the absorbed energy terms and are strictly comparable.

However, if the main aim is the assessment of energy absorption capacity inherent in the

numerical modelling (more specifically the energy dissipation), the influence of the input

energy can be “removed” by dividing Eqs. (6.16) by their second members. The following

Mu··e re+ Mu··g–=

Mu··a ra+ Mu··g–=

u u··

r

Mu··e( )T ued0

t

∫ reT ued

0

t

∫+ Mu··g–( )T ued0

t

∫=

Mu··a( )T uad0

t

∫ raT uad

0

t

∫+ Mu··g–( )T uad0

t

∫=

u· eTMu· e( ) 2⁄ u· a

TMu· a( ) 2⁄

Eek Ea

k

Eea Ea

a

Eek Ee

a+ Eei=

Eak Ea

a+ Eai=

Eei Ea

i

Eea

Eaa Ee

i Eai

Eea Ea

a

Page 347: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 323

equations are obtained

(6.17)

where and refer to the relative absorbed energy and, similarly,

and are the relative kinetic energies. Since these last terms vanish when velocities approach

zero, it follows that terms tend to 1; moreover, and contain the elastic (recoverable)

and the dissipated (irrecoverable) contributions, for which the curves range between 1 and

the minimum envelope representing the relative dissipated energy.

The input energy diagrams for both for the 0.4S7 and the 1.5S7 tests were obtained and are

included in Figure 6.37.

Figure 6.37 Total input energy for experimental and numerical analysis

For the latter test, experimental and analytical diagrams do compare quite well as a result of

the good agreement of storey displacements, particularly during the time interval with larger

ground accelerations. Deviations of displacements after 5.0s time (see Figure 6.35-a)) did not

significantly affect the input energy comparison due to the lower ground accelerations.

On the contrary, for the 0.4S7 test, rather different input energy diagrams are obtained, as is

apparent in Figure 6.37-a). Note, however, that after 4.0s time, the experimental and analytical

curves are similar, but shifted by an approximately constant amount of energy. Such energy

εek εe

a+ 1=

εak εa

a+ 1=

εea Ee

a Eei⁄= εa

a Eaa Ea

i⁄= εek

εak

εa εea εa

a

εa

b) 1.5S7 testa) 0.4S7 testTime (s)

.0 .5

1.0 1.5 2.0 2.5 3.0

3.5 4.0 4.5 5.0

ENERGY (kJ)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 .0 2.5

5.0 7.5

10.0 12.5 15.0

17.5 20.0

22.5 25.0

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 Time (s)

ENERGY (x1.E2 kJ)

Analytic

Experimental

Page 348: Seismic Assessment of Reinforced Concrete Frame Structures

324 Chapter 6

deficit in the numerical analysis, arises from the underestimation of displacements (caused by

overevaluation of stiffness) during the time interval with larger ground accelerations (note the

displacement deviations in Figure 6.30-a) and the significant accelerations in Figure 5.4-a),

between 2.0s and 4.0s time).

The minimum envelopes of relative absorbed energy diagrams are included in Figure 6.38, for

the tests under analysis. As stated before, such curves represent the relative dissipated energy,

which in the 0.4S7 test includes both hysteretic and viscous components. In particular, for this

test the curve of energy dissipated only by hysteresis is also included (see Figure 6.38-a)).

Figure 6.38 Relative absorbed energy for experimental and numerical analysis

Besides the need of the viscous damping to match the energy response of the 0.4S7 test (as

already concluded before), it can be seen that analytical relative dissipated energy approxi-

mately follows the experimental one, once the major deviations of displacements are over-

come. In fact, during the initial time interval, results do not agree, but, by the end of the

analysis, the fraction of dissipated energy is almost the same in both cases, in spite of the sig-

nificantly different input energy.

For the 1.5S7 test the fraction of energy dissipated by hysteresis in the analysis globally fol-

lows quite well the experimental one and, in the end, again very similar values are obtained in

both cases. This fact was expectable in view of the good agreement between total absorbed

energy diagrams (see Figure 6.36-d)) and between the input energy ones (see Figure 6.37-b)).

This allows to conclude that, in spite of displacement responses somewhat different from the

Time (s)

Relative ENERGY

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 .00 .11

.22 .33

.44 .55 .66

.77 .88

.99 1.10

Analytic

Experimental

Analytic (hysteretic only)

Time (s)

Relative ENERGY

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 .00 .11

.22 .33

.44 .55 .66

.77 .88

.99 1.10

b) 1.5S7 testa) 0.4S7 test

Page 349: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 325

experimental ones, the numerical modelling gives good predictions for the dissipated energy in

the dynamic analysis of the bare frame tests.

6.4.4.5 Summary of dynamic analysis results

The most relevant aspects about the dynamic simulation of the tests are briefly summarized in

the following paragraphs.

Structural frequencies obtained by the analysis of bare frame seismic tests do compare well

with the experimental ones, in spite of the initial frequency deviations due to the uncertainty on

modelling the actual initial stiffness.

The simulation of the 0.4S7 test has shown good adequacy of the model to follow the dynamic

response in the post-cracking range. Besides the initial deviations related with the initial stiff-

ness estimate, the response is very well simulated (both in frequency and in amplitude) by

including viscous damping forces characterized by the experimentally measured factor (1.8%),

equivalent to the hysteretic dissipation due to cracking.

In turn, for the 1.5S7 test, the response is very reasonably reproduced by the model, particu-

larly if no viscous damping is included (actually, as adopted in the experiment). Indeed, for the

involved deformation level (in the post-yielding range), the behaviour is mainly controlled by

the steel, whose dissipation characteristics in the non-linear range, are not likely to be simu-

lated by the viscous damping obtained for the initial undamaged state of the structure. Addi-

tionally, the hysteretic model rules are set up accounting for the steel behaviour, particularly its

dissipation properties. In this context it can be accepted that, more than useless, the inclusion

of viscous damping for post-yielding range simulations may lead to poor approximations of the

actual behaviour, if hysteretic dissipation is inherent in the model.

Finally, note that, where viscous damping was included in the foregoing analyses, its contribu-

tion was quite significant, in spite of the low factor of 1.8%. In the pre-yielding range, due to

the low hysteretic dissipation of the model, that contribution reached over 60% of the total dis-

sipated energy, while for post-yielding behaviour it was about 30%. Therefore, should viscous

damping be necessary to account for energy dissipation, its participation can be expected to

reach high levels, even for low viscous damping factors. Indeed, similar conclusions had been

drawn in previous works by Uang and Bertero (1988).

Page 350: Seismic Assessment of Reinforced Concrete Frame Structures

326 Chapter 6

6.4.5 Flexibility element versus fixed length plastic hinge (F.H.) modelling

The assessment of the flexibility model performance, when compared to other modelling

options, should focus on quality of results, computation efficiency and modelling simplicity

from the user point-of-view. The first issue is duly addressed in this section, but no reliable and

general conclusion can be drawn for the remaining ones, because the features of the flexibility

and the F.H. models are not strictly comparable as far as discretization is concerned (this topic

is further discussed below).

Some global results (storey displacements, inter-storey drifts and shear forces, energy dia-

grams, etc.) obtained by the F.H. model are presented along with the experimental ones, in

order to compare both flexibility and F.H. model ability to trace out the experimental response.

For this purpose, only the 0.4S7 and the 1.5S7 tests were simulated, through both static and

dynamic analysis.

6.4.5.1 Assumptions for F.H. modelling

Structural discretization for F.H. modelling was done by using beam-column elements already

available in CASTEM2000, namely:

• A two-node element with one Gauss point, thus assuming uniform bending moment and

curvature, including shear deformation according to Timoshenko formulation and support-

ing a non-linear moment-curvature behaviour law as described in Appendix B; this element

will be simply referred as the TIMO element.

• The classical Bernoulli two-node beam-element with linear elastic behaviour.

Since no global element was available in CASTEM2000 including both plastic hinge and rigid

lengths at member end zones, modelling of each beam and column had to be done by means of

the association of several elements as shown in Figure 6.39.

Figure 6.39 Member discretization for fixed length plastic hinge (F.H.) analysis

Two rigid elements were considered to account for the joint region and modelled with “infi-

1 2

Linear elastic element

lp1 lp2

Non-Linear element

lr1 lr2

Rigid element

Page 351: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 327

nite” stiffness Bernoulli elements; according to the flexibility discretization, only for the

beams were considered these rigid elements.

Lengths of plastic hinge zones, each one modelled by one single TIMO element, were taken

and , respectively for beams and columns, where h stands for

the full-depth of the cross-section; smaller plastic hinge lengths were adopted for the columns

because less plastic development is expected there. Note that such option is precisely one diffi-

cult issue of the F.H. modelling, because hinge lengths depend on the expected deformation

level and their distribution must be known or assumed a priori to provide non-linear elements

in the adequate locations.

The elastic characteristics and model parameters for the non-linear elements were taken almost

the same as those adopted in the flexibility formulation for the corresponding member end sec-

tions. The only exception refers to the strength degradation parameter , adopted with signifi-

cantly lower values than those referred in 6.2.4, due to the location of the common point which

differs from that considered in the flexibility formulation model.

Actually, for the original trilinear model as described in Appendix B, the common points for

unloading branches exist in the elastic stiffness line 1 (see Figure 3.15), whereas for the modi-

fied model they lie in the fully-cracked stiffness lines 2 or 3; this means that, if the same value

of is used for the original model as for the modified one, less stiffness degradation and more

energy dissipation are obtained (due to “fatter” loops). Therefore, from the observation of the

shear-drift diagrams the storey initial stiffness were estimated using the 0.4S7 test results and,

both yielding thresholds and hypothetical common points for storey unloading stiffness were

roughly obtained from the 1.5S7 test diagrams. The order of magnitude of was estimated

and, upon several trials, a uniform value of 0.2 led to reasonable agreement of the dissipated

energy diagrams between analysis and experiment. However, it is recognized that such an

energy related criterion is not robust enough for the estimation of because other effects are

involved, such as the pinched shape of diagrams and the global stiffness.

For the internal linear element, the elastic and geometric section characteristics were taken as

the average of each end section.

Vertical loads, uniformly distributed along the flexible length, were also included by means of

their nodal equivalents.

lp1lp2

h= = lp1lp2

h 2⁄= =

α

α

α

α

Page 352: Seismic Assessment of Reinforced Concrete Frame Structures

328 Chapter 6

The adopted F.H. discretization, consisting of 5 and 3 elements per beam and column, respec-

tively, actually prevents any comparison of model efficiency because the inherent increase of

total number of elements and nodes requires much more computation time. In fact this was

confirmed by the CPU time for the F.H. calculation much higher than for flexibility computa-

tion; for the static simulations of 0.4S7 and 1.5S7 tests, 18 min and 20 min, respectively, were

required with the flexibility formulation, while with F.H. modelling 60 min and 89 min were

necessary to perform 400 steps of analysis.

Finally, concerning modelling simplicity, it is obvious that the present flexibility formulation

renders the discretization and data preparation tasks much easier than the F.H. model subdivi-

sion of each member in several elements. However, it is recognized that such increased sim-

plicity arises from features, such as the rigid and the plastic hinge lengths being automatically

included in only one element, which are not the main added value of the present flexibility for-

mulation; indeed they could be easily included in a F.H. formulation by means of condensation

of d.o.f. as adopted by Coelho (1992).

6.4.5.2 Discussion on F.H. modelling and comparison with flexibility analysis results

Static analysis

Results from the static analysis with the F.H. formulation are shown in Figures 6.40 and 6.41,

respectively for the 0.4S7 and 1.5S7 tests.

Concerning the 0.4S7 test, the most relevant aspect is the overestimation of storey shear, as

apparent from Figure 6.40-a) referring to the first storey. Until 2.0s, the response is similar to

that of the flexibility formulation (see Figure 6.16-a) for comparison), but then shear peaks

become visibly overestimated due to the higher stiffness involved. Clearly, the linear behav-

iour of the internal elements, with constant uncracked stiffness, prevents the F.H. analytical

response to approach the experimental one (see the shear-drift diagram in Figure 6.40-b)), con-

trarily to the results of flexibility formulation. This effect occurs in the whole structure and can

be confirmed by the generalized overestimation of storey peak shear shown in Figure 6.40-d).

However, the lack of energy dissipation inherent in the flexibility analysis in the pre-yielding

range, is partially overcome by the F.H. modelling, because the model allows for residual

deformation immediately after cracking (see Figure 6.40-b)). Consequently, the dissipated

Page 353: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 329

energy becomes better simulated, as is clearly apparent from Figure 6.40-c) when compared

with Figure 6.16-c), in spite of the overestimated stiffness.

Figure 6.40 0.4S7 test. Static analysis with F.H. modelling

Results from the 1.5S7 test show a response quite similar to the flexibility modelling one,

which can be confirmed by comparing Figure 6.41-a) with Figure 6.18-a). Again, up to 2.0s

time, the first storey shear is overestimated due to the high stiffness involved (see the shear-

drift diagram in Figure 6.41-b)), after which results do agree very well with experimental ones,

as obtained previously for the flexibility analysis. Note that the main difference between the

Time (s)

Time (s)

DRIFT (x1.E-3)

Storey 1

b) 1st storey shear-drift diagram a) 1st storey shear

c) Evolution of 1st storey energy

SHEAR (kN)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 -7.5

-6.0

-4.5

-3.0

-1.5

.0

1.5

3.0

4.5

6.0

7.5 x1.E2

-5.0 -4.0 -3.0 -2.0 -1.0 .0 1.0 2.0 3.0 4.0 5.0 -7.5

-6.0

-4.5

-3.0

-1.5

.0

1.5

3.0

4.5

6.0

7.5 SHEAR (kN)x1.E2

ENERGY (kJ)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0 .0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

10.0 STOREY

.0 1.5 3.0 4.5 6.0 7.5

1

2

3

4

SHEAR (x1.E2 kN)

d) Profile of storey peak shear

Analytic

Experimental

Page 354: Seismic Assessment of Reinforced Concrete Frame Structures

330 Chapter 6

two models is related to the pre-yielding behaviour; once the response enters in the post-yield-

ing range, both models have almost identical features and, therefore, the corresponding

responses are not likely to deviate. In the unloading phases, slight differences may be found

due to the adaptations of the degradation rule for unloading stiffness.

Figure 6.41 1.5S7 test. Static analysis with F.H. modelling

The initial “fatter” loops induce higher energy dissipation, but after 2.0s time the energy dia-

gram develops quite similarly to that of the flexibility analysis shown in Figure 6.18-c); the

final energy value is approximately the same, revealing that the initial energy “jump” is com-

pensated by slightly less dissipation during unloading phases.

Time (s)

Time (s)

DRIFT (%)

Storey 1

b) 1st storey shear-drift diagram a) 1st storey shear

c) Evolution of 1st storey energy

SHEAR (kN)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0

x1.E3 SHEAR (kN)x1.E3

ENERGY (*1.E2 kJ)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0

STOREY

1

2

3

4

SHEAR (x1.E3 kN)

d) Profile of storey peak shear

Analytic

Experimental

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

-2.5 -2.0 -1.5 -1.0 -.5 .0 .5 1.0 1.5 2.0 2.5

.0

.2

.4

.6

.8

1.0

1.2

1.4

1.6

1.8

2.0

.0 .3 .6 .9 1.2 1.5

Page 355: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 331

Peak values of storey shear are larger than those obtained from flexibility analysis, particularly

in the three upper storeys, which may be related with the higher stiffness in the present model-

ling. However, in absolute terms, peak shear deviation from the experimental values do not

significantly differ from those of the flexibility analysis.

The overall comparison of results, obtained from static analysis with both the flexibility and

the F.H. modelling of the 0.4S7 and 1.5S7 tests, allows to conclude that the experimental

response is better estimated by the flexibility formulation, particularly in the pre-yielding

range where the progressive adaptation of member stiffness is quite important. Better approxi-

mation of storey shear is achieved, although for energy assessment purposes the F.H. formula-

tion reveals itself more adequate before yielding. However, after yielding development,

structural responses provided by both modelling strategies do not differ significantly.

Dynamic analysis

For dynamic analysis purposes, model parameters for the F.H. formulation were kept as for the

static analyses. Concerning viscous damping, the option made was to select the parameters

leading to the best results with the flexibility approach, namely 1.8% and 0%, respectively for

the 0.4S7 and 1.5S7 tests. However, as before, this aspect turned out to be critical for the good

assessment of the dynamic response.

The results of dynamic analysis for 0.4S7 are shown in Figure 6.42: displacement time histo-

ries of the first and fourth stories (Figures 6.42-a) and b)) reveal poorer approximation to the

experimental ones, when compared against the flexibility results in Figures 6.30-a). In spite of

a reasonable estimate of the top displacement, the response clearly deviates from the experi-

mental time history: right after the positive peak near 2.2s time, when significant cracking

occurred, the response by F.H. modelling stays behind the flexibility analysis due to the higher

stiffness of the linear internal elements; in the subsequent peaks the response becomes progres-

sively out-of-phase and over-damped after the highest peak.

Such over-damping effect suggests that the adopted viscous damping might not be adequate

simultaneously with the hysteretic dissipation provided by this model, mostly if one bears in

mind the quite good results obtained with the same viscous damping and no hysteretic dissipa-

tion in the cracking phase, as adopted in the flexibility formulation.

Page 356: Seismic Assessment of Reinforced Concrete Frame Structures

332 Chapter 6

Figure 6.42 0.4S7 test. Dynamic analysis with F.H. modelling and 1.8% viscous damping

Concerning the inter-storey drift (see Figure 6.42-c)), peak values are clearly worse estimated

in the two upper storeys and better evaluated in the first one, comparatively to the flexibility

modelling results (Figure 6.31-b)); however, the maximum peak drift (in the second storey) is

almost the same in both analyses. In turn, the peak storey shear shows that, where the drift

develops more (i.e., in the three lower storeys), the force response is quite overestimated due to

the excessive stiffness involved.

In order to check the influence of viscous damping, a similar calculation was performed but

Storey 1

d) Profile of storey peak shearb) 1st storey displacement

a) 4th storey displacement

STOREY

.0 1.0 2.0 3.0 4.0 5.0

1

2

3

4

DRIFT (x1.E-3)

c) Profile of storey peak-drift

STOREY

1

2

3

4

SHEAR (x1.E2 kN)

Analytic

Experimental

Time (s)

DISPL. (m)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0

x1.E-2 Storey 4

-5.0

-4.0

-3.0

-2.0

-1.0

.0

1.0

2.0

3.0

4.0

5.0

Time (s)

DISPL. (m)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0

x1.E-2

-5.0

-4.0

-3.0

-2.0

-1.0

.0

1.0

2.0

3.0

4.0

5.0

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0

Page 357: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 333

with no viscous damping. A clear amplification of the response is obtained, as apparent from

the top displacement time history and the inter-storey drift peak profile shown in Figure 6.43.

Figure 6.43 0.4S7 test. Dynamic analysis with F.H. modelling and zero viscous damping

In spite of the hysteretic dissipation inherent in the F.H. modelling, the response amplification

due to viscous damping “removal” is much more evident than it was for the flexibility model-

ling (for comparison check Figure 6.30-a) against Figure 6.32-a) for the flexibility analyses,

and Figure 6.42-a) against Figure 6.43-a) for the F.H. calculations). This fact suggests that,

beyond the problem of adopting an adequate viscous damping factor, the length of zones where

hysteretic dissipation is taken into account may be responsible for the cruder agreement

between F.H. modelling and the experimental results. Indeed, this modelling allows dissipation

only in the plastic hinges, while the cracked zones develop further outside those lengths (as

shown in the cracking pattern included in Figure 6.17-c)), which means that dissipation may

not be sufficiently accounted for in the assumed non-linear zones.

Figure 6.44 shows the 1.5S7 test results, as previously referred, for no viscous damping. Com-

paring these results with those obtained by the flexibility formulation (see Figures 6.35-a) and

c)), the peak displacements and drifts become fairly amplified in the F.H. modelling.

Observing the top displacement response and the first storey shear-drift diagram, the first two

cycles are characterized by displacements lower than the experimental ones, due to the higher

Storey 4

b) Profile of storey peak drifta) 4th storey displacement

STOREY

1

2

3

4

DRIFT (x1.E-3)

Analytic

Experimental

Time (s)

DISPL. (m)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0

x1.E-2

-5.0

-4.0

-3.0

-2.0

-1.0

.0

1.0

2.0

3.0

4.0

5.0

.0 1.0 2.0 3.0 4.0 5.0

Page 358: Seismic Assessment of Reinforced Concrete Frame Structures

334 Chapter 6

stiffness of the structure. Consequently, the dissipated energy in those cycles is very low and,

as a determinant factor for the subsequent cycles, this contributes to push the following peak

deformations to higher levels in order to compensate for the low energy dissipation.

Figure 6.44 1.5S7 test. Dynamic analysis with F.H. modelling and zero viscous damping

In terms of peak storey shear, results do not significantly differ from the flexibility analysis;

the slight increase, which can be observed upon comparison of Figure 6.44-d) and Figure 6.36-

c), arises from the larger development of deformations in the post-yielding range taking place

in the F.H. modelling.

Time (s) DRIFT (%)

Storey 4

b) 1st storey shear-drift diagram a) 4th storey displacement

c) Profile of storey peak drift

DISPL. (m)

.0 .8 1.6 2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0

SHEAR (kN)x1.E3

SHEAR (x1.E3 kN)

d) Profile of storey peak shear

Analytic

Experimental

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

.0 .3 .6 .9 1.2 1.5

-.30

-.24

-.18

-.12

-.06

.00

.06

.12

.18

.24

.30

STOREY

.0 0.7 1.4 2.1 2.8 3.5

1

2

3

4

DRIFT (%)

STOREY

1

2

3

4

Storey 1

-3.5 -2.8 -2.1 -1.4 -.7 .0 .7 1.4 2.1 2.8 3.5

Page 359: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 335

6.4.5.3 Summary of F.H. and flexibility modelling comparison

In the previous section it has been found that F.H. modelling gives less good results than the

flexibility element model in the pre-yielding range of static analysis, while in the post-yielding

analysis both modelling strategies led to similar responses; the adaptive stiffness of the mem-

ber has been considered to play the most important role.

In dynamic analysis, the viscous damping selection appears a more difficult problem because

the F.H. model itself accounts for dissipation in the pre-yielding range and, therefore, both

damping sources contribute to dissipate energy. In turn, the consideration of zero viscous

damping in F.H. modelling led to overestimates of the response, because the dissipation on

actually cracked zones is not duly taken into account; additionally, in the post-yielding range

the simulation is less good than the flexibility one due to the high stiffness of the internal linear

elastic elements.

Note that better results could be expected if the stiffness of the internal elements was reduced,

but that is for sure one of the most difficult options of the F.H. modelling. First of all, one has

to have an a priori estimate of the expected deformation in order to define where and how

much the stiffness has to be reduced, which is neither easy nor straightforward. Then, if a

given analysis is likely to go through a wide range of deformation levels, a certain stiffness

reduction may be adequate up to a certain limit, but not valid from then on.

The above topic resumes the main advantage of the flexibility formulation over the traditional

fixed length plastic hinge modelling: the progressive adaptation of the member stiffness,

throughout the deformation range, allows to closely follow the overall structure stiffness and,

therefore, the vibration frequencies as determinant factors for adequate assessment of the seis-

mic response.

Finally, concerning the dissipative characteristics, it is worth mentioning that, despite the low

energy dissipation obtained with the flexibility model in the pre-yielding range, results indicate

good predictions of the experimental behaviour if the fully-cracked stiffness is assumed pro-

gressively developing along structural members in association with viscous damping forces (of

Rayleigh type) characterized by a factor estimated for the initial state of the structure. After

yielding has started, the viscous damping is deemed unnecessary as long as the structural stiff-

ness modifications are adequately traced. Note that modifying viscous damping characteristics

Page 360: Seismic Assessment of Reinforced Concrete Frame Structures

336 Chapter 6

depending on the deformation range, is still an inconvenient procedure for compensating a

model drawback, but it is surely less troublesome than any attempt to modify the structural

stiffness of a F.H. based model during the analysis.

6.5 Conclusions

The present chapter has focused on the numerical analyses of the four-storey full-scale build-

ing by means of the flexibility element model. In this context, all the modelling assumptions

were first described and the necessary data was defined based on the actual structure layout

and the average properties available from constituent material tests; then, both the skeleton

curves and the hysteretic behaviour parameters for global section modelling were defined.

Where damping forces needed to be explicitly included, they have been considered of viscous

type and characterized by the classical Rayleigh damping matrix proportional to the mass and

stiffness matrices. The adequacy of including explicit damping forces has been discussed in

view of the model features concerning energy dissipation by hysteresis.

Infill panels were modelled by diagonal struts according to a previous work by Combescure

(1996) on the analysis of the infilled configurations of this structure. The basic steps of such

work have been briefly recalled to address the background studies (refined non-linear analysis

and application to experimentally tested specimens) on which the diagonal strut model was

based, particularly its validation and derivation of model parameters. The difficulties of defin-

ing panel characteristics, viz those related to the initial stiffness and the panel-frame interface,

have been pointed out as they clearly affect the numerical simulations of tests.

Damage indices for the quantification of structural damage have been presented, based on the

most relevant proposals available in the literature. The need to incorporate the combined

effects of large strain excursions and several repetitions of load reversal in a unique damage

index has been highlighted. Among the proposals fulfilling this requirement, the damage index

adopted herein was chosen based on an assessment study of damage indices by Fardis et al.

(1993), according to which the well known Park and Ang index (Park et al. (1987a)) provides

one of the best agreements with experimental results. This index was considered here in terms

of chord rotations at each element end section, coincident with the element rotations in the

space without rigid body modes. The quantification of element related parameters for damage

Page 361: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 337

evaluation (viz, the ultimate rotation , the dissipated hysteretic energy and the respec-

tive factor) was carefully discussed in view of the available proposals and the specific fea-

tures of the element.

A wide set of numerical analyses was performed in order to closely trace out the structure

behaviour throughout the several stages and to systematically compare it with the experimental

results.

Although without experimental counterpart, a static pushover analysis was first performed by

applying an inverted triangular distribution of storey forces to the structure, from which the

global yielding displacement and the maximum base shear were estimated. The former

appeared somewhat overevaluated while the base shear became slightly underestimated in

relation to the approximate experimental values based on seismic tests; this might be due to the

low post-yielding stiffness of the global section model and, eventually, due to strength mecha-

nisms differently activated in the pushover analysis and in the tests. However, the roughness of

the experimental estimates may also justify the discrepancy.

Static simulations of the performed tests were carried out by applying the experimentally

imposed displacements, in order to assess the modelling results without involving dynamic

effects.

The simulations of the bare frame seismic tests has shown quite good results for stabilized

cracked behaviour, although with some problems to account for energy dissipation due to the

origin-oriented features of the section model in the cracked range. Once yielding sets in (as for

the high level test), good agreement is still found between numerical and experimental results.

Energy dissipation becomes better captured, although slightly overestimated due to difficulties

in simulating the pinching effect arising from anchorage slippage. The overall distribution of

the damage index confirms the expected dissipation mechanism (strong column - weak beam)

and the maximum values (0.4) conform with the observed state of the structure.

Numerical static analyses for the infilled frame seismic tests provided poor simulations of the

experimental behaviour due to the estimates of model parameters for infill panels (particularly

the initial stiffness). However, simulations became better after and where significant cracking

occurred in infills, because the response became controlled by the reinforced concrete frame.

θu Ed∫β

Page 362: Seismic Assessment of Reinforced Concrete Frame Structures

338 Chapter 6

Results from the final cyclic test simulations were difficult to compare with experimental ones

due to the repair of damaged zones; nevertheless, the global strength was reasonably captured

(though slightly underestimated). However, the most relevant aspect is the difficulty in simu-

lating the unloading stiffness degradation for large ductility levels, which, for the present

model rule leads to over-degradation and, consequently, to lower energy dissipation.

In the context of dynamic analysis, the assessment of structural frequencies has shown good

agreement of numerical and experimental results for the bare frame structure.

Only the bare structure low and high level seismic tests were numerically simulated, having

shown good adequacy of the model to trace out the dynamic response in the post-cracking

range. However, viscous damping forces had to be included to compensate for the lack of hys-

teretic dissipation in the model, and the initially measured viscous damping factor (1.8%)

proved to yield very good results while the response is dominated by cracking. In turn, for the

high level test, when the response became mainly controlled by the steel behaviour, the viscous

damping became useless (even inadequate) since the dissipation features are taken into account

in the section model; therefore, the simulation of this test with zero viscous damping led to

very reasonable estimates of experimental results, concerning both displacement and energy

dissipation.

Whether or not the viscous damping should be included, its contribution to energy dissipation

has been found to reach significant levels, even for low factors. This means that viscous damp-

ing inclusion should be carefully judged upon the specific dissipative features of the model

before and after yielding.

Some numerical simulations with the flexibility element model were compared with analyses

by traditional fixed length plastic hinge modelling. The comparison has shown that flexibility

modelling leads to dynamic analysis results closer to the experimental ones, which arises from

its main advantage over traditional fixed hinge modelling, i.e., the progressive adaptation of

the member stiffness, throughout the deformation range. Since the overall structural stiffness

can be closely followed, so are the vibration frequencies and a better simulation of the dynamic

behaviour can be obtained.

Additionally, this comparison has emphasized that, as long as structural modifications are con-

veniently traced out (as with the flexibility modelling), good predictions of the experimental

Page 363: Seismic Assessment of Reinforced Concrete Frame Structures

ANALYSIS OF THE 4-STOREY FULL-SCALE BUILDING 339

behaviour can be obtained by adopting fully cracked stiffness along with viscous damping

forces in the pre-yielding range and by removing viscous dissipation in the post-yielding

behaviour.

Overall, the flexibility element modelling has shown good adequacy to simulate the seismic

behaviour of the structure under analysis. However, the following issues are recognized to

require further development: the element dissipation features in the pre-yielding range, the

unloading stiffness degradation, the pinching effect and the rebar slippage simulations. The

first is directly related with the flexibility element formulation proposed herein, while the sec-

ond and the third depend only of the global section model; the last issue is better accounted for

by specific elements as proposed by Filippou et al. (1992) or Monti et al. (1993).

Page 364: Seismic Assessment of Reinforced Concrete Frame Structures

340 Chapter 6

Page 365: Seismic Assessment of Reinforced Concrete Frame Structures

Chapter 7

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES

DESIGNED ACCORDING TO EC8

7.1 Introduction

The Eurocode 8 is included in the set of nine Eurocodes presently being discussed and pub-

lished as prestandards defining the common rules for structural and geotechnical design. The

EC8 is concerned with the seismic design and, due to its innovative features relatively to exist-

ing national codes, it is considered an advanced code and even “the latest word in codified

earthquake resistant design” (Carvalho et al. (1996)).

Aiming at the safe, yet economic, design of earthquake resistant structures, the innovative fea-

tures of EC8 (e.g. the establishment of a serviceability limit state for damage limitation under

seismic action, or the adoption of capacity design procedures for reinforced concrete building

structures) have been subjected to a european-wide discussion and testing process over the last

few years, in order to gather relevant information for code improvement or revision to become

a definitive european standard.

Aiming at EC8 testing and validation, a unified effort has been carried out within the frame-

work of the pioneering research project entitled “Prenormative Research in support of Euroc-

ode 8” (PREC8), co-funded by the European Commission and National Authorities.

The present chapter includes the presentation and discussion of results from the numerical seis-

mic analyses of RC frame structures performed under the PREC8 programme, as a result of

our activity at the Joint Research Centre (JRC). The major concern is in line with the project

entitled “Reinforced concrete frames and walls” which included the design of trial cases (con-

Page 366: Seismic Assessment of Reinforced Concrete Frame Structures

342 Chapter 7

stituting a set of typical RC buildings) and the evaluation of their non-linear seismic response,

in order to find out the implications of EC8 provisions on the seismic behaviour of building

structures. Some trial cases, viz those having the so called basic configurations 2 and 6, were

analysed by the author within the JRC team activities and the results were extensively reported

in Arêde et al. (1996). Some modifications and improvements were subsequently performed

which are already included in the results recalled in this chapter.

For completeness, a general overview of the PREC8 project is provided in 7.2, specifically

focusing on reinforced concrete frame and wall structures. Then, the building configurations 2

and 6 are presented in 7.3 where details are included concerning structural layout, trial cases,

loads, modelling assumptions and response variables to express the analysis results. The most

important results obtained from the non-linear analyses are described and discussed in 7.4,

bearing in mind the expected seismic performance of structures designed according to the

basic philosophy underlying EC8. Finally, the main conclusions of the present chapter are

summarized in 7.5, particularly focusing on the assessment of the structures analysed herein.

7.2 The PREC8 project

7.2.1 Basics of EC8

According to EC8, the basic principles of safe seismic design require that in case of seismic

events: a) human life protection is ensured, b) damage is limited and c) important structures for

civil protection (lifelines, hospitals, etc.) are kept operational. This is assumed to be accom-

plished if the structural design ensures the two following basic requirements to be fulfilled:

• no collapse of the structure under the design earthquake (having a very small probability of

occurrence) which stands for the performance assessment referring to the Ultimate Limit

State (ULS)

• the damage is kept within repairable limits when the structure is subjected to a seismic

action with a larger probability of occurrence, thus assuring the structural performance with

reference to the Serviceability Limit State (SLS)

The seismic action is represented in EC8 by the elastic response spectrum, based on the design

ground acceleration (indeed, the effective peak ground acceleration in rock or firm soil) and on

other parameters conveying the influence of local ground conditions (sub-soil classes). The

Page 367: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 343

design ground acceleration is assumed, for EC8 purposes, to approximately represent the seis-

mic hazard variable in a given zone. The seismic zonation is to be defined by National Author-

ities and the hazard is assumed to be constant within each seismic zone, such that the design

ground acceleration typically corresponds to a reference return period of 475 years (equivalent

to 10% exceedance probability in 50 years).

The ULS verifications rely on the exploitation of the structural ductile capacity, which means

that a non-linear ductile response is generally assumed to take place in order to dissipate the

seismic input energy; however, the limiting case of non-dissipative structures is also consid-

ered in EC8. For design purposes the non-linear ductile response is approximately taken into

account by means of the concept of a global behaviour factor, through which a design spectrum

suitable for linear analysis methods is derived from the elastic response spectrum; the design

spectrum is thus used to characterize the design seismic action for ULS verifications.

Instead, the SLS verifications are carried out for a seismic action derived from the design one

by a reduction factor taking into account the lower return period of earthquakes associated with

that limit state.

The structural regularity issue is treated separately in terms of regularity in plan (mainly affect-

ing the requirements on structural analysis models) and of regularity in elevation (which also

influences the behaviour factor values, more reduced for irregular structures). For the more

general cases, EC8 requirements may lead to significant computational demands, such as those

related to spatial dynamic analysis where accidental eccentricities and simultaneous action of

earthquake components are to be included.

The energy dissipation issue for RC structures is particularly developed in EC8 by the adoption

of different Ductility Classes (DC), reflecting a certain trade-off between structural strength

and ductility, indeed, the two basic and simultaneous requirements in earthquake resistant

structures. According to EC8, three ductility classes can be considered: a) Ductility Class L

(DCL) corresponding to structures designed and detailed according to EC2, although supple-

mented with additional detailing rules to enhance the available ductility; b) Ductility Class M

(DCM) to which correspond structures designed, dimensioned and detailed enabling the struc-

tural behaviour to enter well within the inelastic range without brittle failures and c) Ductility

Class H (DCH) corresponding to structures whose design, dimensioning and detailing provi-

Page 368: Seismic Assessment of Reinforced Concrete Frame Structures

344 Chapter 7

sions ensure the development of chosen stable mechanisms allowing large hysteretic energy

dissipation.

In operational terms, ductility classes are distinguished by different behaviour factors (q),

reducing as the ductility class decreases, according to the proportionality rule of 1.00, 0.75 and

0.50, respectively for DCH, DCM and DCL. Moreover, for the DCM and DCH, Capacity

Design procedures (Paulay and Priestley (1992), Eurocode 8 (1994)) are adopted in order to

provide the structure with a more suitable mechanism for energy dissipation. Typically, for

reinforced concrete frame structures, Capacity Design aims at allowing inelastic excursions

only at beam end regions and at the base of ground floor columns, such that, despite the dam-

age occurred in those regions, the overall stability of the structure is still assured under gravity

load actions. To this end, a certain sequence of individual member design has to be followed in

order to account for the actual strength of adjacent members.

Although the practical application of EC8 as a provisional norm has been already undertaken,

some priority needs of research support were identified and a broad scientific network was set

up to accomplish a Pre-normative Research project in support of EC8 (PREC8) (Pinto and

Calvi (1996)). In view of different types of structures with specific behaviour features, distinct

topics were covered, namely: a) RC frames and walls; b) Infilled frames; c) Bridges and d)

Foundations and retaining walls.

The major concern in the present chapter focuses on the first topic, whose main scope was to

investigate “the interrelation between a number of design parameters used in EC8, which in a

combined form, influence the non-linear behaviour of structures subjected to earthquake

motions” (Carvalho et al. (1996), Pinto and Calvi (1996)). These parameters are the regularity

classification, the analysis methods, the behaviour factor values and the effects of capacity

design procedures. An additional issue which has drawn the attention of researchers aimed at

assessing how safe and economic are the structures designed according to EC8 provisions and

whether equivalent results are obtained by designing for different Ductility Classes. Further

details on this topic of the PREC8 project are included in the next section.

7.2.2 The RC frame structure topic

A large parametric study of RC frame and wall structures was performed in the PREC8 con-

text, particularly aiming at the assessment of ductility classes and respective behaviour factors.

Page 369: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 345

A set of buildings was designed according to Eurocodes 2 and 8, for which other parameters

than the ductility class were also varied, in order to constitute a number of different situations.

Their seismic response was analytically evaluated for earthquakes of increasing intensity, aim-

ing at checking whether the EC8 provisions lead to a satisfactory structural response through-

out the different trial cases.

The set of buildings consisted of 26 distinct RC structures, distributed as follows according to

a specific envisaged purpose (Carvalho et al. (1996)):

• buildings for three types of uses, in order to highlight to which extent the relative impor-

tance of vertical versus horizontal loads may affect the result of capacity design provisions;

• buildings with four different heights (3, 4, 8 and 12 storeys), in order to check the influence

of the natural period of the structure and the effect of minimum design provisions likely to

prevail at the upper storeys of taller buildings;

• buildings located in two distinct seismic zones, characterized by design accelerations 0.15g

and 0.30g (respectively, standing for low/medium and medium/high seismicity), where the

adequacy of code rules should be checked and compared;

• buildings with framed structures and a central core, in order to introduce the effect of cou-

pled frame-wall behaviour and assess the outcome of code provisions for walls.

Buildings for numerical analysis were grouped in six basic configurations (1 to 6) and, with the

exception of configuration 4 (an industrial building) which had three storeys and large spans,

all structures were based on a similar rectangular plan layout with 15 m (3 bays of 5 m) by 20

m (5 bays of 4 m or 3 bays of 8/4/8 m). Storeys were typically 3 m high, except for the indus-

trial building and for configuration 6 where four columns were suppressed at the ground level

while the remaining ones are 4.5 m high (an irregular situation was sought, related to a softer

first storey and to some columns resting on first floor beams). Generally, dynamic analysis

methods were used for the design seismic analysis, but in two specific configurations addi-

tional analysis by static methods was carried out (as allowed by EC8) in order to check their

influence on the seismic performance of structures.

The whole process of building design was carried out at University of Patras (Fardis (1995)) in

a computerized way, from which it became clear that EC8 application requires computer anal-

ysis using appropriate software, particularly due to seismic action combinations and to capac-

ity design procedures, the latter introducing links between several phases of the design.

Page 370: Seismic Assessment of Reinforced Concrete Frame Structures

346 Chapter 7

Additionally, interesting information was gathered concerning the characteristics of EC8

designed structures and, particularly, regarding the required quantities of steel and concrete

depending on the ductility class. It was found that the total quantities of both materials is

approximately the same for the three ductility classes (Fardis (1995), Carvalho et al. (1996)),

although differently distributed: a) the ratio of column-to-beam total steel shifts from about

55%-45% for DCL, to 60%-40% for DCM and to 65%-35% for DCH, and b) the ratio of longi-

tudinal-to-transverse steel varies from about 80%-20% for DCL, to 75%-25% for DCM and to

60%-40% for DCH. These trends on the steel distributions arise from the capacity design pro-

visions of EC8, not foreseen for DCL but gaining increasing importance as the ductility class

increases.

Once the design results were available, the non-linear analysis of the 26 buildings was carried

out by six partners of the PREC8 project. Aiming at a certain control of results, each trial case

was analysed by two different teams and some basic assumptions were commonly adopted by

all partners (e.g. mean values of resistances, input accelerograms and damping quantification).

The analyses were performed separately for each horizontal direction with four distinct artifi-

cial accelerograms generated to fit the EC8 spectrum for soil type B and for peak ground accel-

erations of 0.15g or 0.30g, whichever matches the design acceleration. Each trial case was

analysed for the nominal intensity (1.0 times the peak acceleration) and also for two other

increased intensities (with scaling factors of 1.5 and 2.0) aiming at roughly tracing the vulner-

ability functions of the structures.

The JRC team was commissioned to perform the non-linear analyses of 9 trial cases with the

basic configurations 2 and 6, consisting of 8 storey frame buildings, regular in the cases of con-

figuration 2 and irregular in the other cases, which are further described in the next section.

7.3 The building configurations 2 and 6

7.3.1 General comments and structure layout

The general layout of the two basic configurations 2 and 6 is illustrated in Figure 7.1, both in

plan and in elevation, where the global coordinate system is also included.

Page 371: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 347

3.00

24.00

4.00

3.00 3.00 3.00 3.00 3.00 3.00 3.00

4.00

4.00

4.00

4.00

Y

X

4.00

4.00

4.00

4.00

4.00

Col

umns

rem

oved

in

the

grou

nd fl

oor

5.005.005.00

Y

X

3.00

25.50

3.00 3.00 3.00 3.00 3.00 3.00 4.50

5.005.005.00

Figu

re 7

.1B

asic

con

figur

atio

ns o

f the

eig

ht s

tore

y tr

ial c

ases

(PR

EC8)

a) C

onfig

urat

ion

2b)

Con

figur

atio

n 6

Page 372: Seismic Assessment of Reinforced Concrete Frame Structures

348 Chapter 7

The structures are symmetric in both horizontal directions (XX and YY) and, while configura-

tion 2 is regular in plan and in elevation, configuration 6 exhibits two sources of irregularity in

elevation: a) the first storey appears to be softer than the remaining ones, due to its greater

height and to the absence of some columns cut off at the first storey; b) the existence of these

cut-off columns, supported by medium-long span beams, may generate itself additional

demands which can spread all over the structure.

Each configuration was designed according to EC2 and EC8, for different ductility classes (L,

M and H) and for two design accelerations (0.15g and 0.3g). Furthermore, the case of configu-

ration 6 for ductility class M and design acceleration 0.3g was also designed using the simpli-

fied static analysis of paragraph 3.3.2 of EC8, Part 1.2, herein labelled as “Mst”. The

combination of these design assumptions (ductility class, design acceleration and analysis

method) leads to the nine distinct trial cases listed in Table 7.1, which also includes the design

behaviour factors and the reference names (labels) identifying each trial case in the following

sections.

All the relevant data obtained from the design process is extensively described in Fardis

(1994), namely concerning section design forces, cross-section dimensions, reinforcement

details (longitudinal and transversal) and adopted slab widths contributing for beam strength

and stiffness; however, for completeness purposes, the cross-section dimensions of trial case

members are summarized in Table 7.2. Columns have uniform cross-sections in elevation and,

for configuration 6, the columns removed in the ground floor are referred to as “Cut-off” col-

umns. Except for the first storey beams of configuration 6 (which are more robust to support

the cut-off columns) the beam cross-sections are also uniform in elevation and almost uniform

in plan.

Table 7.1 Trial cases, design behaviour factors and earthquake intensities

Config. Design Accel.(a)

Ductility Class

Reference Name

Behaviour Factor (q)

Earthquake Intensities (*a)

20.15g

L 2_15L 2.5 - 1.0 1.5 2.0M 2_15M 3.75 - 1.0 1.5 2.0

0.30g M 2_30M 3.75 - 1.0 1.5 2.0H 2_30H 5.0 0.5 1.0 1.5 2.0

6

0.15g L 6_15L 2.0 - 1.0 1.5 2.0M 6_15M 3.0 - 1.0 1.5 2.0

0.30gM 6_30M 3.0 - 1.0 1.5 2.0

Mst 6_30Mst 3.0 - 1.0 1.5 2.0H 6_30H 4.0 0.5 1.0 1.5 2.0

Page 373: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 349

7.3.2 Vertical static loads and seismic action

According to the data used in the design, the following vertical static loads, per unit area, have

been considered:

• slab self-weight: ws = 3.5 kN/m2

• finishing: wf = 2.0 kN/m2

• live load: ql = 2.0 kN/m2

Using the appropriate combination coefficients for variable actions as prescribed in EC8, the

static vertical loads to be adopted simultaneously with the seismic action have been computed

by the following rules:

• top floor: ws + wf + 0.30*ql

• other floors: ws + wf + 0.15*ql

Vertical loads uniformly distributed per unit length in beams were obtained from the previous

values with the adequate influence areas and the beam self-weight was also included. These

loads were applied prior to any seismic input in order to start the seismic analysis with the

effects of dead and live loads already taken into account (namely, in what concerns stiffness).

The seismic action was simulated by four artificial accelerograms, shown in Figure 7.2, that

were provided to all the participant teams in the PREC8 project.

Table 7.2 Member cross-sectional dimensions (m)

CaseColumns Beams (b/h)

Internal External Corner Dir. X Dir. X(long span)

Dir. Y

2_15L .60x.60 .55x.55 .50x.50 .30x.60 .30x.60 .30x.602_15M .60x.60 .55x.55 .50x.50 .25x.50 .30x.60 25x.502_30M .70x.70 .60x.60 .60x.60 .30x.60 .30x.60 .30x.602_30H .70x.70 .60x.60 .60x.60 .30x.60 .30x.60 .30x.60

Internal External Cut-off Dir. X(1st floor)

Dir. X(2-8th floors)

Dir. Y

6_15L .70x.70 .60x.60 .50x.30 .30x.80 .30x.60 .30x.606_15M .70x.70 .60x.60 .50x.30 .30x.80 .30x.60 .30x.606_30M .80x.80 .70x.70 .50x.30 .30x.80 .30x.60 .30x.606_30Mst .80x.80 .70x.70 .50x.30 .30x.80 .30x.60 .30x.606_30H .80x.80 .70x.70 .50x.30 .30x.80 .30x.60 .30x.60

Page 374: Seismic Assessment of Reinforced Concrete Frame Structures

350 Chapter 7

Figure 7.2 Artificial accelerograms (S1...S4) and response spectra (5% damping) fitting the

EC8 response spectrum

Time [s]

Acceleration (m/s2)

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 -1.0 -.8 -.6 -.4 -.2 .0 .2 .4 .6 .8 1.0 Max.: 0.96

Time [s]

Acceleration (m/s2)

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 -1.0 -.8 -.6 -.4 -.2 .0 .2 .4 .6 .8 1.0 Max.: 1.0

Time [s]

Acceleration (m/s2)

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 -1.0 -.8 -.6 -.4 -.2 .0 .2 .4 .6 .8 1.0 Max.: 1.06

Time [s]

Acceleration (m/s2)

.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 -1.0 -.8 -.6 -.4 -.2 .0 .2 .4 .6 .8 1.0 Max. : 0.84

Period [s]

(m/s2)

.0 .5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 .0

.5

1.0

1.5

2.0

2.5

3.0

Period [s]

(m/s2)

.0 .5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 .0

.5

1.0

1.5

2.0

2.5

3.0

Period [s]

(m/s2)

.0 .5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 .0

.5

1.0

1.5

2.0

2.5

3.0

Period [s]

(m/s2)

.0 .5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 .0

.5

1.0

1.5

2.0

2.5

3.0

EC8 - Soil B

S1

S2

S3

S4

Page 375: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 351

The accelerograms of 10 s duration were generated to fit the EC8 response spectrum for soil

type B and 5% damping and normalized to a unitary base acceleration. The set of accelero-

grams was scaled for the design acceleration corresponding to each of the nine trial cases listed

in Table 7.1 and then factored by the intensities included in the same table. Thus, considering

all the trial cases, with all the earthquake intensities, for all accelerograms, in both directions

XX and YY independently, a total of 232 non-linear dynamic analyses was performed.

7.3.3 Structure modelling

7.3.3.1 Discretization

The frame structures shown in Figure 7.1 were discretized for independent planar analyses in

the XX and YY directions. Taking profit of the symmetry, only the association of two distinct

frames (one internal and other external) is considered in each direction of analysis with double

values of stiffness, vertical static load and mass. The corresponding structural systems are

illustrated in Figure 7.3 for both configurations 2 and 6 in the two directions of analysis.

Equal horizontal displacements were imposed to all the nodes at the same floor in order to

accomplish the assumption of rigid floor diaphragm. Each structural member was discretized

by only one flexibility global element as adopted for the analysis of the four storey building

tested at ELSA (see Chapter 6). Despite the problems found in simulating the experimental

response of the ELSA building for high ductility levels (mainly due to the modelling of

unloading stiffness degradation) the model was still adopted for the present chapter analyses

because the foreseen intensities are not expected to develop such high ductility levels as those

attained for the final cyclic tests. Thus, even with the referred modelling limitation, the model

is likely to reproduce reasonably well the structural behaviour for the expected ductility range.

Vertical static loads were considered by the approximation inherent in the flexibility element

model, i.e., the distributed load was lumped into equivalent concentrated forces at end and

mid-span sections.

Different characteristics were assigned to the left and right parts of each element, according to

the actual data of the corresponding member end section. However, these characteristics are

assumed uniform along each element part, meaning that details of span reinforcement or varia-

tion of slab width are not taken into account.

Page 376: Seismic Assessment of Reinforced Concrete Frame Structures

352 Chapter 7

Rigid lengths were assumed only for the beam ends (with half width of the nearest column)

and the effect of pull-out of reinforcement bars from the joint was not considered. Shear

behaviour was considered linear elastic for both beams and columns, using shear reduced areas

equal to the cross-sectional area divided by 1.2 (rectangular cross-section).

Figure 7.3 Structural systems of planar frame associations

7.3.3.2 Mass, damping and natural frequencies

The structural mass has been computed for each floor according to the vertical static loads

referred to in 7.3.2, and equally distributed by the nodes belonging to that floor. Table 7.3 sum-

marizes the total mass values for each floor of all the trial cases. The total mass of the building

is included for comparison with the value used in the design (also listed in the same table),

which shows small deviations of no more than 5%. Note that the contribution of half inter-sto-

rey height of columns is also included in the floor mass, which explains the different values for

the top floor and also for the first floor of configuration 6.

Internal Frame External Frame

X

Z

Storey level

8 7

2 1

Direction XX Direction YY

Internal Frame External Frame

a) Configuration 2

Internal Frame External Frame

Direction XX Direction YY

Internal Frame External Frame

b) Configuration 6

6 5 4 3

Storey level

8 7

2 1

6 5 4 3

X

Z

Page 377: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 353

The first and second natural frequencies, in both directions of analysis, are given in Table 7.4

for all cases assumed with uncracked behaviour, including also the frequency values obtained

in the design process (Fardis (1994)). The values shown here differ slightly from the design

ones, with a general overestimation trend that agrees with the lower mass values and that may

be also related with little differences of slab width. The general agreement of frequencies is

quite acceptable and can be regarded as an “extra” global check of the structure input data.

The global inspection of the first mode frequency values (as obtained here) shows that, with

the exception of the 2_15M, 6_15L and 6_15M structures, for most cases in the XX direction,

the fundamental vibration period “falls” into the “constant” acceleration branch of the spec-

trum (check Figure 7.2) although very close to the transition point for the descending branch

(at 0.6s period or 1.67 Hz frequency). Instead, for the YY direction, the structures are more

flexible and, most of them, fall into the descending branch.

Table 7.3 Floor masses. Adopted values and design values (in brackets)

CaseMass per Floor (ton)

Total (ton)Floor 1 Floors 2 - 7 Floor 8

2_15L 270 270 255 2145 (2236)2_15M 260 260 244 2064 (2080)2_30M 284 284 262 2250 (2333)2_30H 284 284 262 2250 (2333)6_15L 296 272 256 2184 (2299)6_15M 296 272 256 2184 (2299)6_30M 316 288 263 2307 (2328)6_30Mst 316 288 263 2307 (2328)6_30H 316 288 263 2307 (2328)

Table 7.4 Frequencies (Hz) for all cases (design values in brackets)

Case1st Mode Frequency 2nd Mode Frequency

XX YY XX YY2_15L 1.71 (1.64) 1.49 (1.45) 5.20 (5.00) 4.53 (4.45)2_15M 1.52 (1.52) 1.23 (1.23) 4.67 (4.69) 3.79 (3.86)2_30M 1.86 (1.78) 1.63 (1.58) 5.69 (5.48) 4.98 (4.91)2_30H 1.85 (1.78) 1.61 (1.58) 5.66 (5.48) 4.94 (4.91)6_15L 1.63 (1.52) 1.50 (1.40) 4.91 (4.61) 4.58 (4.33)6_15M 1.61 (1.52) 1.47 (1.40) 4.86 (4.61) 4.53 (4.33)6_30M 1.80 (1.71) 1.68 (1.59) 5.45 (5.26) 5.14 (4.97)6_30Mst 1.83 (1.71) 1.72 (1.59) 5.52 (5.26) 5.21 (4.97)6_30H 1.77 (1.71) 1.63 (1.59) 5.88 (5.26) 5.05 (4.97)

Page 378: Seismic Assessment of Reinforced Concrete Frame Structures

354 Chapter 7

This fact is quite relevant concerning the structural seismic response because it is responsible

for part of the structural overstrength. Indeed, for structures having vibration periods near or in

the descending branch of the spectrum, the structure softening due to concrete cracking and

reinforcement yielding, increases the vibration period, which leads to seismic forces lower

than the design ones and, consequently, to lower demands in the structure comparatively to

those foreseen in the design. This constitutes the so called “demand-side” overstrength (Fardis

and Panagiotakos (1997)), which is different, in nature, from the “supply-side” overstrength

(typically related with the difference between the design values of material properties and the

mean values as used in non-linear analyses) and leads to a reduction of ductility demands.

Viscous damping was included, proportional to the mass and stiffness (Rayleigh type), with a

damping factor of 2% for both the first and the second modes, as adopted for some analyses of

the four-storey ELSA building. Despite the conclusions drawn in 6.4.4.5, pointing out that vis-

cous damping might not be adequate for hysteretic response simulation in the post-yielding

range, the factor of 2% was still kept for the present analyses because it corresponds to a fixed

data value imposed to all partners participating in the PREC8 project.

7.3.3.3 Moment-curvature constitutive relations for global section behaviour

The trilinear moment-curvature primary curves for global section modelling were obtained by

the process described in 4.2.3, for which the material properties were adopted according to typ-

ical mean values of the concrete class (C25/30) and the steel grade (B500 Tempcore) as used in

the design. Thus, according to the notation of stress-strain diagrams shown in 4.2.3.2, the fol-

lowing values were considered:

• Steel: fsy = 585 MPa; Es = 200 GPa; Esh = 1.7 GPa; εsm = 0.090;

• Concrete: fc0 = 33 MPa; εc0 = 0.0022; fct = 2.6 MPa; Ec = 30.5 GPa;

For the concrete model, the final comments of 6.2.2.2 apply: the confinement parameters were

obtained from the cross-section and transverse reinforcement data and residual stresses were

taken as 0% and 20% of the peak stress, respectively, for unconfined and confined concrete.

For M-ϕ curve evaluation, the effective slab width participating in beam strength and stiffness

was considered using values available in the design information (Fardis (1994)). Concerning

the axial force, null values were taken for the beams, while constant values due the vertical

static loads were considered for columns.

Page 379: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 355

By contrast with the ELSA building, no fibre refinement was done for more accurate definition

of yielding and of post-yielding in columns (as described in 6.2.3). However, special care was

taken with the ultimate point definition for column sections where the presence of axial load

can lead to a rather curved shape of the post-yielding branch of the M-ϕ diagram. Actually, in

some cases, the ultimate point can lie on the softening part of the curve, as schematically

shown in Figure 7.4, in such a low level that the approximated post-yielding line YU has a neg-

ative slope. In order to avoid this situation an extra (intermediate) point I is determined and the

line YI is adopted for the post-yielding branch. Additionally, a modified ultimate point U’ is

considered such that the strain energy between Y and U’ is the same as that between Y and U.

The intermediate point I was taken here as the supplementary point referred to in 4.2.3.4, cor-

responding to an average strain in the most compressed concrete fibre or in the most tensioned

reinforcement between strain values at yielding and at ultimate conditions (see Figure 4.16).

Figure 7.4 Approximation of the post-yielding branch of the (M-ϕ) diagrams in columns

The set of M-ϕ curves for all the different cross-sections of configuration 6, in the direction

YY, is included in Figures 7.5 and 7.6 both for beams and columns, respectively. Although no

detailed information can be extracted from those figures for a specific cross section, they are

useful to provide an idea of the order of magnitude of the resistances involved and also of the

available curvature ductility (at least in terms of a rough estimate). The following aspects can

be pointed out from those figures:

• Ultimate curvatures for beams in positive bending (ϕ+u) are much more uniform than those

in negative bending (ϕ-u), which is due to the fact that for ϕ+

u the bottom steel layer (having

always a smaller area than the top one) is tensioned and the compression zone can spread

along the effective slab width. Thus, no concrete crushing is likely to occur for this bending

ϕ

My

Mc

M

ϕc ϕy

C

Y

Mu

ϕu

U′

U

I

Page 380: Seismic Assessment of Reinforced Concrete Frame Structures

356 Chapter 7

direction (positive), while for negative bending large compressions develop in the web bot-

tom fibres due to the high top steel content. It is apparent that this effect is much more criti-

cal for the lower ductility classes (i.e., DCL for 0.15g and DCM for 0.30g) because in such

cases higher design seismic forces are involved.

• Yielding moments (My) of beams decrease when the ductility class level is increased; more-

over, as expected between two ductility classes (e.g. M and L), the ratio of corresponding

yielding moments is not far from the inverse ratio of behaviour factors

. The relation between these two ratios depends on practical design reasons

(detailing) and on the relative magnitude of vertical static loads compared to the seismic

forces, but still the value of is helpful to check the trend of changes in My between duc-

tility classes. Similar reasons explain the fact that between the 0.30g and the 0.15g design

options the corresponding ratio of My values is not 2.0 as could be expected. Furthermore, it

is worth mentioning that, for the static design option 30Mst of configuration 6, the values of

My are higher than for the case 30M which is related to the simplified static procedure used

for the seismic analysis that usually overestimates the action effects.

• For columns, the ultimate curvature values and their regularity, as well as the corresponding

ductility factor, are higher for increasing ductility class levels and the same happens for the

uniformity of the post-yielding stiffness.

• The trend of My variations in columns when ductility class is varied is not as clear as it is

for beams, because, apart from the behaviour factors ratio, the capacity design magnifica-

tion factors play an important role in column design. Actually, these factors are different

from one ductility class to another and, due to its dependency on the design and the resistant

moments distribution, their effect can go either in the same or in the opposite direction of

the ductility class variation. For this reason some cases were found where the increase of

ductility class seems not to modify the column My values (configuration 2, not shown

herein) while in other cases a clear modification is noticeable (configuration 6).

Finally, concerning the hysteretic behaviour parameters, similar options were adopted as for

the four-storey ELSA building (see 6.2.4). For the unloading stiffness degradation the parame-

ter was considered for both positive and negative directions, the pinching effect was

taken into account only due to top and bottom reinforcement asymmetry ( factors were

obtained by Eq. (B.3) in Appendix B) and the strength degradation factors were calculated

using Eq. (B.4).

rm MyM My

L⁄=( )

rq qM qL⁄( )1–

=

rq

α 4.0=

γ

Page 381: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 357

6_30

M_Y

6_15

M_Y

-.2

5 -.

20 -

.15

-.10

-.0

5 .

00

.05

.10

.1

5 .

20

.25

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

(x1.

e3)

6_15

L_Y

6_30

H_Y

6_30

Mst

_Y

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

-.2

5 -.

20 -

.15

-.10

-.0

5 .

00

.05

.10

.1

5 .

20

.25

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

(x1.

e3)

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

-.2

5 -.

20 -

.15

-.10

-.0

5 .

00

.05

.10

.1

5 .

20

.25

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

(x1.

e3)

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

-.2

5 -.

20 -

.15

-.10

-.0

5 .

00

.05

.10

.1

5 .

20

.25

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

(x1.

e3)

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

-.2

5 -.

20 -

.15

-.10

-.0

5 .

00

.05

.10

.1

5 .

20

.25

-1.5

-1.2

-.9

-.6

-.3

.0

.3

.6

.9

1.2

1.5

(x1.

e3)

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

Figu

re 7

.5B

eam

sec

tion

mom

ent-c

urva

ture

dia

gram

s of

all

case

s w

ith c

onfig

urat

ion

6 in

dire

ctio

n YY

M

Page 382: Seismic Assessment of Reinforced Concrete Frame Structures

358 Chapter 7

.0

0 .

02

.04

.06

.0

8 .

10

.12

.14

.1

6 .

18

.20

.0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

.0

0 .

02

.04

.06

.0

8 .

10

.12

.14

.1

6 .

18

.20

.0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

.0

0 .

02

.04

.06

.0

8 .

10

.12

.14

.1

6 .

18

.20

.0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

.00

.02

.0

4 .

06

.08

.10

.1

2 .

14

.16

.18

.2

0

.0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

.0

0 .

02

.04

.06

.0

8 .

10

.12

.14

.1

6 .

18

.20

.0

.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

6_30

M_Y

6_15

M_Y

(x

1.e3

)6_

15L

_Y

6_30

H_Y

6_30

Mst

_Y

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

(x1.

e3)

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

(x1.

e3)

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

(x1.

e3)

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

(x1.

e3)

Mom

ent (

kN.m

)

Cur

vatu

re (m

-1)

Figu

re 7

.6C

olum

n se

ctio

n m

omen

t-cur

vatu

re d

iagr

ams

of a

ll ca

ses

with

con

figur

atio

n 6

in d

irect

ion

YY

Page 383: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 359

7.4 Non-linear seismic analysis of building configurations 2 and 6

7.4.1 General

The results from the non-linear analysis of the structures are presented and discussed in the

present section. They are expressed by common response variables such as total base shear, top

displacement (or total drift), inter-storey drift (relative to the storey height), member displace-

ments, ductility factors and damage indices.

Both ductility and damage refer to member end chord rotations and are defined as for the 4-sto-

rey ELSA building, following the procedures and expressions described in 6.3.2.

Generally, results refer to average values calculated from the response to the four accelero-

grams considered; typically they are categorized according to the basic design assumptions

(design acceleration and ductility class) and to the action intensity.

7.4.2 Structural strength

The structural strength engaged during seismic response is an important issue to be addressed

since it can provide an insight into the global behaviour. The total peak base shear provides a

measure of the global strength and can be compared to the design forces based on which the

member dimensioning and detailing were carried out. Such forces (available from the design

information in Fardis (1994)) are given in Table 7.5 as a fraction of the total weight, showing

that, except for the configuration 2 designed for 0.30g, the design seismic forces almost coin-

cide in both directions XX and YY.

The ratio of peak base shear from the non-linear seismic analysis to the design seismic

forces is plotted in Figure 7.7 for all trial cases and intensity levels. This ratio, henceforth

Table 7.5 Design base shear force ratio to structure weight (seismic coefficient)

Case Configuration 2 Configuration 6XX YY XX YY

15L 0.138 0.134 0.169 0.16315M 0.084 0.086 0.113 0.10930M 0.190 0.205 0.233 0.23230H 0.145 0.162 0.175 0.17630Mst -- -- 0.249 0.249

Rseismmax( )

Rd( )

Page 384: Seismic Assessment of Reinforced Concrete Frame Structures

360 Chapter 7

denoted by , is a measure of the structure overstrength for loading conditions exceeding the

design seismic force and it can be seen that maximum values of about 2.0 can be found for

twice the design intensity.

Figure 7.7 Global overstrength

The major sources contributing for that overstrength can be categorized as follows:

• the difference between mean and design values of constitutive material strengths; mean val-

ues as used in non-linear analysis are approximately 1.3 and 2.0 times the design values,

respectively for steel and concrete;

• minimum reinforcement requirements and the unavoidable bar rounding up;

• capacity design requirements and gravity load rather than earthquake dominated design of

some sections;

• the post-yielding hardening at the section level;

• the fact that different strength mechanisms may be activated in the design and in the analy-

sis, such that, for a seismic action with the design intensity, only part of the critical zones

have yielded while the remaining ones keep behaving in the pre-yielding range (non-simul-

taneous yielding of all the critical zones that are assumed in the design).

ψm

a) Configuration 2 b) Configuration 6

Direction YY

Direction XX

0.0

0.5

1.0

1.5

2.0

2.5

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.5

1.0

1.5

2.0

2.5

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Global OverStrength Global OverStrength

0.0

0.5

1.0

1.5

2.0

2.5

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.5

1.0

1.5

2.0

2.5

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

ψm Rseismmax Rd⁄=( )

Page 385: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 361

The first three sources typically affect the level of the global yielding threshold, while the two

last ones mostly contribute for the smoothness of the yielding transition and for the hardening

in the post-yielding range. Although all these overstrength sources are mixed up in the global

response, their influence on the yielding level can be approximately assessed by means of a

push-over analysis consisting of an inverted triangular distribution of lateral forces monotoni-

cally applied to the structures (as also performed for the 4-storey ELSA building). Thus, base

shear - top displacement diagrams were obtained for all trial cases in both directions of analy-

sis, and, despite the difficulty of defining what the global yielding point is, approximate esti-

mates were determined for the yielding base shear force according to the criterion described

next; the base shear - top displacement curve for one trial case is shown in Figure 7.8.

Figure 7.8 Base shear - top displacement curve for the C2_15L case, direction XX. Definition

of global yielding force

Push-over analysis curves were considered up to a top displacement approximately corre-

sponding to a total drift of 2%, which is an upper bound well above the maximum drift found

in the analysis (about 1.5% for twice the design intensity); such a displacement level is deemed

to correspond to a stabilized hardening range of the global response. A fictitious yielding point

Y* is defined by the intersection of straight lines approximating the pre- and post-yielding

curve zones and the global yielding point Y is taken lying on the curve and having the same

displacement as Y*; thus, the yielding base shear Ry is read at the level of point Y.

The ratio of Ry to Rd (design force), standing for the overstrength at yielding and denoted by

TOP DISPL.(m)

BASE SHEAR (kN)

.00 .05 .10 .15 .20 .25 .30 .35 .40 .45 .50 .00

.60

1.20

1.80

2.40

3.00

3.60

4.20

4.80

5.40

6.00 (x1.E3)

Y*

YRy

Page 386: Seismic Assessment of Reinforced Concrete Frame Structures

362 Chapter 7

, is given in Table 7.6 and shows that a significant reserve of strength is found for most

cases; values range from about 1.1 to 1.5, but with a clear trend to exceed 1.3. Average values

of are also listed, showing that the ductility class does not appear to influence the over-

strength at yielding, while for the lower design acceleration slightly larger values of are

obtained. This is consistent with the higher relative weight of gravity loads in the 0.15g cases,

whose design is more likely to be controlled by other load combinations than the seismic ones

and, therefore, to have enhanced strength to lateral loads. Finally, it is emphasized that the glo-

bal average factor is about 1.36, thus agreeing with the above referred overstrength factor

(1.3) for steel that mostly controls the yielding threshold.

The comparison of the force with Ry allows to check, in a global sense, whether the

post-yielding behaviour has been actually entered; the ratio of to Ry, herein denoted by

and designated by global hardening factor, is plotted in Figure 7.9 for all trial cases.

The factor is actually a measure of the structural hardening at the global level for the cases

where , but it also means that for cases having clearly below 1.0, no yielding has

taken place. Indeed, that is the case of the DCL structures for the design intensity, due to their

important reserve of strength and to the fact that just a low portion of seismic action is to be

absorbed by ductile behaviour (q-factors of 2.5 and 2.0); this justifies also the fact that for

twice the design intensity, the peak base shear stays below 1.2 times the yielding one, while for

the other trial cases, the hardening factor reaches values as high as 1.5.

The analysis of is also valuable to understand the global overstrength factor for the

design intensity (see Figure 7.7), whose average value is about 1.0 for DCL structures. In these

cases, an elastic response to the design intensity earthquakes would generate average base-

shear forces about q (2.5 and 2.0) times the design ones; since no yielding was found, it fol-

Table 7.6 Overstrength factors at yielding

CaseConfiguration 2 Configuration 6XX YY XX YY Average

15L 1.50 1.29 1.41 1.35 1.3915M 1.50 1.15 1.49 1.36 1.3830M 1.38 1.16 1.36 1.31 1.3030H 1.38 1.09 1.45 1.29 1.3030Mst -- -- 1.52 1.49 1.50Average 1.44 1.17 1.45 1.36 1.36

ψo

ψo

ψo

ψo

ψo Ry Rd⁄=( )

Rseismmax

Rseismmax

ψh

ψh

ψh 1> ψh

ψh ψm( )

Page 387: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 363

lows that the obtained elastic force reduction to the design force level must arise from crack-

ing. Actually, the corresponding frequency drop induces a reduction of the spectral ordinate in

the hyperbolic descending branch of the elastic spectrum (see Figure 7.2), which for the DCL

structures appear to be about 40% to 50% of the original values (based on the uncracked

behaviour as assumed in the design); indeed, similar results have been reported by Fardis and

Panagiotakos (1997) for other structures within the PREC8 framework, reflecting the impor-

tant reduction of seismic demands due to structural cracked behaviour.

Figure 7.9 Global hardening factor

For structures of higher ductility classes (M and H), the higher q factors (3 to 5) lead to force

reductions (compared to the elastic ones) larger than those induced by the frequency drop due

to cracking, which means that after the cracking effect, there is still a portion of seismic

demand to be absorbed by post-yielding behaviour. This is confirmed in Figure 7.9 for almost

all DCM and DCH cases, whose factors for the design intensity exceed 1.0, except for the

6_15M structures.

From the above presented results, the most relevant topics can be summarized as follows:

• the important strength reserve, particularly at the yielding level, induces structures designed

a) Configuration 2 b) Configuration 6

Direction YY

Direction XX

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1 x Design

1.5 x Design

2 x Design.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1 x Design

1.5 x Design

2 x Design.

Hardenning Factor Hardenning Factor

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1 x Design

1.5 x Design

2 x Design.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1 x Design

1.5 x Design

2 x Design.

ψh Rseismmax Ry⁄=( )

ψh

Page 388: Seismic Assessment of Reinforced Concrete Frame Structures

364 Chapter 7

for medium seismicity (0.15g) to behave in the pre-yielding range or just at imminent yield-

ing for the design intensity;

• for cases exceeding the global yielding, the global overstrength factor ranges between 1.3

and 1.5 for the design intensity, reaching values of about 2.1 for twice the design intensity;

• part of such overstrength is due to global hardening (factors up to 1.5 were found for the

highest intensity) which shows a slight trend to increase with the ductility class level.

7.4.3 Cracking, yielding and damage patterns

In the previous section the importance of structural cracking for the seismic response was high-

lighted. The extension of cracking in structural members is illustrated in Figure 7.10 for the

response of structures with configuration 6, in the direction XX, under the action of the earth-

quake S1 with intensities 1.0 and 2.0.

It is apparent that extensive cracking is found for the design intensity, mostly in beams and

also, to some extent, in the internal columns of the irregular frame. For twice the design inten-

sity, cracking develops further (particularly in the columns) but it is clear that the most signifi-

cant stiffness drop due to cracking takes place for intensity 1.0.

Cracking patterns do not significantly differ for the analysed cases; yet a trend is found for less

cracking development in higher ductility classes (lower seismic forces arising from larger q-

factors) and for larger cracking when the design acceleration is increased from 0.15g to 0.30g.

The rotation ductility patterns are shown in Figure 7.11, while the corresponding damage pat-

terns are included in Figure 7.12 for the same structures. Only positive rotation ductility is

included since the most relevant findings are common to negative rotation ductility patterns;

additionally, only ductility values above 1.0 are plotted as this can provide an indirect and gen-

eral view of the yielding pattern. For comparison between the regular and the irregular config-

urations, Figure 7.13 includes also the damage pattern for configuration 2 in the direction XX.

For the 0.15g designed structures very low ductilities are found for the design intensity as a

consequence of the important strength reserve. However, since these values exceed 1.0, it

appears that local yielding at the element level has developed, while at the global level a pre-

yielding behaviour was found (check values for 6_15L and 6_15M cases in Figure 7.9).

These apparently contradictory findings can be explained by two different reasons:

ψh

Page 389: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 365

• On the one hand, rotation ductility values very close to 1.0 (as in the 6_15L case) may cor-

respond to situations where the plastic hinge has not actually developed. Indeed, at the ele-

ment level, and particularly in beams, the moment distribution along the members may be

more influenced by gravity loads comparatively to the 0.15g earthquake action effects

(which, furthermore, become strongly reduced by cracking development). Consequently,

the inflection point location may significantly deviate from the mid-span section and situa-

tions may occur where the chord rotation exceeds the yielding rotation (for which the

inflection point is assumed at mid-span) without actual development of plastic hinge. Note,

however, that such drawback, inherent in the adopted rotation ductility definition, tends to

vanish for higher intensity of lateral loads.

• On the other hand, the global yielding is not defined at the first onset of yielding in a given

member, which means that, before global yielding, some local hinging may have already

developed. Naturally, as the global response becomes closer to the yielding threshold, this

fact becomes more evident as in the 6_15M case. Indeed, Figure 7.9 shows that this struc-

ture in direction XX reached a maximum base shear closer to the yielding one than the

6_15L case and, accordingly, the corresponding rotation ductility becomes clearly above

1.0.

For twice the design intensity (still for 0.15g designed structures), plastic hinge formation tends

to spread all over the structure, though less evidently in the larger span beams due to their

higher flexibility compared to that of shorter span beams. Except for the cut-off columns (more

slender than the remaining ones), plastic hinging is found only in beams and at the end-zones

of ground floor columns, thus agreeing with the dissipation mechanism foreseen in the design.

Structures designed for 0.30g exhibit larger ductility demands in accordance with the clear

onset of yielding for the design intensity, although somewhat low in view of the adopted

behaviour factor. Note that, for q-factors of 3.0 and 4.0, respectively for 6_30M and 6_30H

structures, ductility demands at the member level could be expected to be at least similar to or

greater than the behaviour factor; however, since a significant overstrength factor is engaged

for the design intensity, the local ductility demand becomes reduced.

Comparing to the 0.15g structures, the 6_30M and 6_30H cases show a more uniform spread

of plastic hinging, which, for intensity 2.0, engages almost all the critical zones of the beam

sidesway mechanism underlying the design philosophy.

Page 390: Seismic Assessment of Reinforced Concrete Frame Structures

366 Chapter 7

Figure 7.10 Cracking pattern: Configuration 6, Direction X under earthquake S1 for intensity 1.0

and 2.0

Intensity: 2.0Intensity: 1.0

6_15L

6_15M

6_30M

6_30H

Case:Internal External Internal External

Page 391: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 367

Figure 7.11 Positive rotation ductility pattern: Configuration 6, Direction X under earthquake S1

for intensity 1.0 and 2.0

Max.= 2.09 Max.= 4.25

Intensity: 2.0Intensity: 1.0

6_15L

6_15M

6_30M

6_30H

Case:

Max.= 1.09 Max.= 2.55

Max.= 1.44 Max.= 2.82

Max.= 2.25 Max.= 4.62

Internal External Internal External

Page 392: Seismic Assessment of Reinforced Concrete Frame Structures

368 Chapter 7

Figure 7.12 Damage pattern: Configuration 6, Direction X under earthquake S1 for intensity 1.0

and 2.0

Intensity: 2.0Intensity: 1.0

6_15L

6_15M

6_30M

6_30H

Case:

Max.= 0.23 Max.= 0.49

Max.= 0.17 Max.= 0.42

Max.= 0.18 Max.= 0.33

Max.= 0.29 Max.= 0.63

Internal External Internal External

Page 393: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 369

Figure 7.13 Damage pattern: Configuration 2, Direction X under earthquake S1 for intensity 1.0

and 2.0

Intensity: 2.0Intensity: 1.0

2_15L

2_15M

2_30M

2_30H

Case:

Max.= 0.30 Max.= 0.59

Max.= 0.23 Max.= 0.40

Max.= 0.15 Max.= 0.37

Max.= 0.27 Max.= 0.65

Internal External Internal External

Page 394: Seismic Assessment of Reinforced Concrete Frame Structures

370 Chapter 7

According to Figures 7.12 and 7.13, damage is found in almost all critical zones, even if no

local yielding has occurred there; this arises from the way how the damage index is calculated,

according to which the deformations before yielding are also taken into account. To some

extent, this can be regarded as a measure of the cracking contribution to the damage in struc-

tural members.

The ratio between maximum damage values for intensities 1.0 and 2.0 approximately follows

the corresponding ratio between rotation ductilities, actually as expected since the damage

index is mostly influenced by the peak rotation values.

The increase of design acceleration leads to larger damage values; on the other hand, for higher

ductility classes (with the same design seismic input) lower damage is obtained as a result of

better design detailing, particularly concerning transversal reinforcement, which enhances the

section (and member) ultimate ductile capacity.

Despite some significant values in the cut-off columns, the damage in configuration 6 (Figure

7.12) is better distributed in the external frame because beam spans are uniform. By contrast,

the large difference of span lengths in the internal frame cause the damage to concentrate in the

shorter central span; a similar result is obtained in the internal frame of configuration 2 (Figure

7.13). Thus, from the obtained results, the damage distribution appears more affected by the

non-uniformity of beam spans rather than the presence of cut-off columns.

Finally, it is noteworthy that higher damage is obtained in the ground floor columns of config-

uration 2 than in configuration 6, for which two reasons can be pointed out. On the one hand,

the column cross-sections of configuration 2 have smaller dimensions than those of configura-

tion 6 and, particularly for the internal frame, for approximately the same axial force; this

results in lower ultimate ductile capacity and, therefore, in higher damage (indeed, the internal

frame shows the most significant increase of base column damage, when passing from irregu-

lar to the regular configuration). On the other hand, configuration 2 has a lower first storey

height, which means that the corresponding drift (and column rotations as well) becomes

increased, thus leading to higher damage than in configuration 6.

7.4.4 Ductility demand and damage distribution in elevation

Besides the global overview of spatial distribution of response variables as given in the previ-

Page 395: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 371

ous section, it is interesting to analyse the peak value distribution of meaningful control varia-

bles along the structure height. For this purpose Figures 7.14 and 7.15 show the elevation

profiles of the maximum column and beam rotation ductilities in each storey, where storey col-

umns are considered those below a given storey level. Similarly, Figures 7.16 and 7.17 include

the corresponding column and beam maximum damage profiles.

From the column ductility profiles (see Figures 7.14) it is apparent that, for the design inten-

sity, ductility values above 1.0 are found only for the 0.30g designed structures, mostly at the

base end-zones of ground floor columns and also in the cut-off columns at their base and top

storey end-zones; however, such ductilities do not exceed 1.5, showing that just incipient

yielding is enforced. Column ductility values are approximately uniform above the 2nd storey

for each ductility class and design intensity case, except for the irregular structure due to the

cut-off columns where the highest demands are concentrated.

Generally, where significant column ductilities above 1.0 are found, the higher ductility class

structures tend to develop larger ductility demands, though this effect is not strictly systematic.

For twice the design intensity, the maximum column ductility for configuration 2 is about 3.0,

occurring in the ground floor columns, whereas for configuration 6 it is about 3.7 in the cut-off

columns.

The beam ductility profiles (see Figures 7.15) confirm that just incipient yielding is obtained

for the design intensity in 0.15g structures; such incipiency is more apparent for DCL struc-

tures where the ductility just slightly exceeds 1.0. For twice the design intensity, the maximum

ductility demand in 0.15g structures is about 3.0, occurring for DCM.

As for the columns, and although not systematic, increased beam ductility demands are

obtained for higher ductility classes, particularly where more significant values are found;

additionally, the maximum demand tends to shift from the upper to the lower floors as the duc-

tility class increases.

In the 0.30g structures, maximum beam ductilities for the design intensity vary between about

2.0 to 3.0, thus confirming the clear onset of yielding; instead, for intensity 2.0, the maximum

demands reach values about 6.2 for the regular structure, whereas in the irregular one this

value slightly exceeds 7.0 in the 2nd storey beams next to the cut-off columns.

Page 396: Seismic Assessment of Reinforced Concrete Frame Structures

372 Chapter 7

Figure 7.14 Column rotation ductility profiles (maxima)

Design Accel.: 0.30gDesign Accel.: 0.15g

Design Accel.: 0.30gDesign Accel.: 0.15g

Direction XX

Direction YY

Direction XX

Direction YY

1

2

3

4

5

6

7

8

0 1 2 3 4 5

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5

Stor

ey

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.MediumDuct.High

1.0 x Des ign Intens.1.5 x Des ign Intens.2.0 x Des ign Intens.Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

Maxim um Colum nDuctility

Maxim um Colum nDuctility

Maxim um Colum nDuctility

Maxim um Colum nDuctility

1

2

3

4

5

6

7

8

0 1 2 3 4 5

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5St

orey

1

2

3

4

5

6

7

8

0 1 2 3 4 5

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5

Stor

ey

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.MediumDuct.High

1.0 x Des ign Intens.1.5 x Des ign Intens.2.0 x Des ign Intens.Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

Maxim um Colum nDuctility

Maxim um Colum nDuctility

Maxim um Colum nDuctility

Maxim um Colum nDuctility

a) Configuration 2

b) Configuration 6

Page 397: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 373

Figure 7.15 Beam rotation ductility profiles (maxima)

Design Accel.: 0.30gDesign Accel.: 0.15g

Design Accel.: 0.30gDesign Accel.: 0.15g

Direction XX

Direction YY

Direction XX

Direction YY

1

2

3

4

5

6

7

8

0 1 2 3 4 5 6 7 8 9

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5 6 7 8 9

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5 6 7 8 9

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5 6 7 8 9

Stor

ey

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

Maxim um BeamDuctility

Maxim um BeamDuctility

Maxim um BeamDuctility

Maxim um BeamDuctility

1

2

3

4

5

6

7

8

0 1 2 3 4 5 6 7 8 9

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5 6 7 8 9

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5 6 7 8 9

Stor

ey

1

2

3

4

5

6

7

8

0 1 2 3 4 5 6 7 8 9

Stor

ey

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

Maxim um BeamDuctility

Maxim um BeamDuctility

Maxim um BeamDuctility

Maxim um BeamDuctility

a) Configuration 2

b) Configuration 6

Page 398: Seismic Assessment of Reinforced Concrete Frame Structures

374 Chapter 7

Figure 7.16 Column damage profiles (maxima)

Design Accel.: 0.30gDesign Accel.: 0.15g

Design Accel.: 0.30gDesign Accel.: 0.15g

Direction XX

Direction YY

Direction XX

Direction YY

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1St

orey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.MediumDuct.High

1.0 x Des ign Intens.1.5 x Des ign Intens.2.0 x Des ign Intens.Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

Maxim um Colum nDam age

Maxim um Colum nDam age

Maxim um Colum nDam age

Maxim um Colum nDam age

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.MediumDuct.High

1.0 x Des ign Intens.1.5 x Des ign Intens.2.0 x Des ign Intens.Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

Maxim um Colum nDam age

Maxim um Colum nDam age

Maxim um Colum nDam age

Maxim um Colum nDam age

a) Configuration 2

b) Configuration 6

Page 399: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 375

Figure 7.17 Beam damage profiles (maxima)

a) Configuration 2

b) Configuration 6

Design Accel.: 0.30gDesign Accel.: 0.15g

Design Accel.: 0.30gDesign Accel.: 0.15g

Direction XX

Direction YY

Direction XX

Direction YY

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.MediumDuct.High

1.0 x Des ign Intens.1.5 x Des ign Intens.2.0 x Des ign Intens.Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

Maxim um BeamDam age

Maxim um BeamDam age

Maxim um BeamDam age

Maxim um BeamDam age

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1

Stor

ey

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.LowDuct.Medium

1.0 x Design Intens .1.5 x Design Intens .2.0 x Design Intens .Duct.MediumDuct.High

1.0 x Des ign Intens.1.5 x Des ign Intens.2.0 x Des ign Intens.Duct.LowDuct.Medium

1.0 x Design Intens.1.5 x Design Intens.2.0 x Design Intens.Duct.MediumDuct.High

Maxim um BeamDam age

Maxim um BeamDam age

Maxim um BeamDam age

Maxim um BeamDam age

Page 400: Seismic Assessment of Reinforced Concrete Frame Structures

376 Chapter 7

Column damage distribution in elevation (see Figures 7.16) essentially follows the trend of the

corresponding ductility demands and it is quite apparent that, for higher ductility classes, less

damage is observed. As already pointed out, larger damage values are found for the base col-

umns of configuration 2 compared to those of configuration 6.

For configuration 2, at the design intensity, the maximum column damage is about 0.25 for the

0.15g structures, whereas for 0.30g cases it just slightly exceeds 0.30; for twice the design

intensity these damage values approximately duplicate. In turn, for configuration 6, the maxi-

mum damage appears to concentrate in the cut-off columns (particularly in the cases designed

for higher seismicity) and for the design intensity, it is about 0.16 and 0.20, respectively for

0.15g and 0.30g cases; again these values approximately duplicate for twice the design inten-

sity.

Beam damage profiles (see Figures 7.17) generally show that higher ductility classes lead to

less damage, although some exceptions are observed for the irregular structure. Still the shift of

peak damage from the upper to the lower storeys is found when the ductility class is increased.

For 0.15g structures the maximum damage for the design intensity is always below 0.2, while

for twice that intensity it is about 0.4. For the 0.30g structures these values appear duplicated,

but the most relevant aspect is the concentration of maximum damage in a couple of storeys,

comparatively to 0.15g structures where the damage is more uniform in elevation (particularly

for configuration 2).

Damage values are typically lower in columns than in beams, following an average reference

ratio of 0.5, which is just violated at base column critical zones where significant damage can

be up to 1.5-2.0 times higher than in the first storey beams. Such high ratio is due to the fact

that maximum damage in beams rarely occurs at the first storey, but still some cases of config-

uration 2 exhibit absolute maximum column damage higher than the beam ones.

7.4.5 Overall analysis of response parameters

Rather than observing the distribution of response variables throughout each structure, the fol-

lowing paragraphs focus on the analysis and comparison of (global) response parameters.

Average estimates of peak values are obtained from the response to the four accelerograms and

the maxima over the entire structure are retained for comparison between trial cases.

Page 401: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 377

The maximum total drift (ratio of top displacement to the structure height) is plotted in Figure

7.18, from which no significant and systematic differences appear between distinct ductility

classes for the same design acceleration; nevertheless, a certain trend can be observed in some

cases for larger drifts when the ductility class is increased, more visible for the 0.15g designed

cases.

In average, configuration 2 leads to drifts higher than configuration 6, as a consequence of the

slender columns of the former; however, even for twice the design intensity, low drift values

are obtained (below 1.6%) when compared to values at the near-failure stage. Indeed, as a ref-

erence, one can look back at the tests on the four storey ELSA building, that reached a total

drift of 4.8% for the final stage (see 5.5.6), to which the failure was considered imminent.

Additionally, it is worth recalling that 1.7% total drift was observed for the high level seismic

test, which, despite a nominal peak ground acceleration of 1.5 times the design one (0.30g),

actually reached almost twice the design acceleration (due to the characteristics of the artifi-

cially generated accelerogram); thus, the drift herein obtained agrees quite well with the exper-

imental evidence on a similar structure type (same design acceleration and ductility class).

Figure 7.18 Total drift (%)

The ratio of total drift at the design intensity to the corresponding design values has been found

a) Configuration 2 b) Configuration 6

Direction YY

Direction XX

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Total Drift (%) Total Drift (%)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Page 402: Seismic Assessment of Reinforced Concrete Frame Structures

378 Chapter 7

(in average) about 1.9, resulting from the structure softening mainly caused by cracking (and

also, to less extent, due to yielding) which, on one hand, tends to amplify displacements and,

on the other hand, reduces the seismic action effects as a consequence of the spectral ordinate

reduction caused by the frequency drop.

Identical comments can be made regarding the maximum inter-storey drift over all storeys as

illustrated in the charts of Figure 7.19. The trend for higher drifts as the ductility class is raised

up becomes more apparent in these charts, but still a wide margin is available with respect to

the inter-storey drifts at failure (note that for the ELSA building values about 7% were

reached). As for the total drift, the regular structure (configuration 2), designed for 0.30g and

DCH as the ELSA building, exhibits about 2.2% inter-storey drift for twice the design inten-

sity, thus approaching the experimentally tested building result of 2.4% drift obtained for the

high level seismic test.

Figure 7.19 Maximum inter-storey drift (%)

Moreover, Eurocode 8 prescribes a serviceability limit state verification (damage control) spe-

cifically concerning the inter-storey drift. According to paragraph 4.3.2 of part 1-2 of EC8, two

limits ((a) and (b)) shall be observed for the inter-storey drift depending on whether or not

“non-structural elements of brittle materials are attached to the structure” that may interfere

a) Configuration 2 b) Configuration 6

Direction YY

Direction XX

0.0

0.5

1.0

1.5

2.0

2.5

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.5

1.0

1.5

2.0

2.5

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Inte r-storey Drift(%) Inter-storey Drift(%)

EC8 (a)EC8 (b)

0.0

0.5

1.0

1.5

2.0

2.5

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.5

1.0

1.5

2.0

2.5

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

EC8 (a)EC8 (b)

Page 403: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 379

with structural deformations. The limits (a) and (b) are, respectively, 0.4*ν % and 0.6*ν %,

where ν stands for the reduction factor of the drift demand to account for the lower seismic

intensity associated with the serviceability limit state. For the present structure category, ν

takes the value 2.0 and the inter-storey drifts for the seismic design intensity are limited by

0.8% and 1.2%, respectively, for conditions (a) and (b). These limits are also indicated in Fig-

ure 7.19 by dashed lines (and denoted by EC8(a) and EC8(b)), showing that, for the design

intensity no case exceeds the limit EC8(b), while most of the 0.30g designed structures do not

fulfil the EC8(a) requirement.

The results of the linear elastic (uncracked) analysis (Fardis (1994)) show that, under design

conditions (say, loading and modelling assumptions) all structures are well within the EC8(a)

limit, which means that the obtained inter-storey drifts shall be compared against 0.8%. It is

apparent that such limit is verified for all the 0.15g designed structures, notwithstanding the

large cracking extent, but for the 0.30g cases it is somewhat exceeded, particularly for configu-

ration 2. This is due to the stiffness drop caused by the generalized spreading of cracking over

the whole structure and confirms the expectable non-conservatism of displacement estimates

based on uncracked behaviour as pointed out in the paragraph 3.1 of part 1-2 of EC8.

Despite this excess of inter-storey drift, the maximum sensitivity coefficient to the second

order effects (as defined in the paragraph 4.2.2 of part 1-2 of EC8) never exceeds the limit 0.1

foreseen in the code, above which the so-called P-∆ effects must be duly accounted for. Indeed,

this is confirmed in Figure 7.20 where the ratio of to the code limit (0.1) is plotted for all

trial cases; although this ratio is, in average, higher for structures of configuration 2 (due to

their relatively larger “slenderness” when compared to those of configuration 6), its maximum

value indicates that the second order overturning moments at a given storey are adequately

below the prescribed limit (i.e., 10% of the storey overturning moments caused by the seismic

inter-storey shear force).

Aiming at the analysis of the seismic action effects at the critical zone level, the maximum

damage over the entire structure is plotted in Figure 7.21 for each trial case and all seismic

intensities.

These results are not easily comparable between different cases because the maximum damage

may occur in distinct locations; note that, as apparent in Figures 7.16 and 7.17, the maximum

θ( )

θ

Page 404: Seismic Assessment of Reinforced Concrete Frame Structures

380 Chapter 7

base column damage is almost as high as the beam damage for some cases. However, still a

quite visible trend is found in Figure 7.21 for lower damage when the ductility class is

increased, indeed as already pointed out for the damage profiles. Additionally, for most cases,

the maximum damage obtained for the irregular structures (configuration 6) is lower than for

structures of configuration 2; this aspect essentially follows the trend already found for the

inter-storey drift (see Figure 7.19) and is mainly related with the relative magnitude of

demands which are higher in configuration 2 due to its less robust members (particularly col-

umns).

Figure 7.20 Sensitivity coefficient

For the design intensity, the maximum damage does not exceed 0.25 for 0.15g designed struc-

tures, while for the 0.30g cases about 0.35 maximum damage occurs in DCM structures that

appear to be the most critical. Particularly for the 0.30g cases, the damage extent is significant,

though within an acceptable level, having in mind the structure performance requirements

under the design intensity earthquakes.

For comparison purposes, it is recalled the damage found for the four storey ELSA building

under a nominal ground acceleration of 1.5 times the design one (0.30g). Among the present

trial cases, the 0.30g_H structure with configuration 2 in direction YY is the one resembling

a) Configuration 2 b) Configuration 6

Direction YY

Direction XX

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Sens itivity Coeff. / Design Value Sens itivity Coeff. / Design Value

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Page 405: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 381

the most with the ELSA structure, i.e. having similar beam spans and storey heights and iden-

tical design acceleration, ductility class and behaviour factor; the corresponding maximum

damage (between 0.36 and 0.53, respectively for intensities 1.5 and 2.0) does not significantly

deviate from the referred ELSA building results, for which a maximum value of 0.41 was

obtained.

Instead, for twice the design intensity, quite important damage develops: almost 0.75 maxi-

mum damage is found for the 0.30g_M structure with configuration 2 in the XX direction. It is

noteworthy that, despite the fact that somewhat larger inter-storey drifts are obtained for most

0.30g_H cases, comparatively to DCM ones, the higher ductility structures tend to develop less

damage by virtue of their enhanced ultimate ductile capacity.

Figure 7.21 Maximum damage

As stated before, some of the above findings do not appear systematic because some maximum

values under comparison do not refer to the same critical zone. Thus, in order to “remove” this

location dependency, the peak damage can be compared in terms of average values over the

entire structure. Such values are often taken as energy weighted averages (Coelho (1992),

Fardis et al. (1993)) and the same criterion was considered herein, adopting the relative energy

contribution of each critical zone to the total energy dissipated in the structure as the weighting

a) Configuration 2 b) Configuration 6

Direction YY

Direction XX

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Maxim um Dam age Maxim um Dam age

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Page 406: Seismic Assessment of Reinforced Concrete Frame Structures

382 Chapter 7

factor. The average damage results are shown in Figure 7.22, where the trend for lower damage

associated with higher ductility class becomes more apparent and systematic.

Figure 7.22 Global (average) damage

Typically, the 0.15g structures exhibit low average damage, even for twice the design intensity,

which is mainly a consequence of their significant overstrength and demand reduction due to

stiffness and frequency decrease. However, among these structures, the DCL ones tend to yield

higher values than the DCM cases for intensity 2.0, because no stringent provisions for ductil-

ity enhancement are required for DCL as they are for DCM.

For the 0.30g structures the average damage slightly exceeds 0.45 in the DCM case of config-

uration 2 (direction XX), which is confirmed to be the most critical. Such average damage is

quite acceptable, particularly because it refers to twice the design intensity. Thus, despite some

locally higher damage, this result highlights the significant reserve of structural capacity to

withstand earthquake action beyond the design one while keeping “its structural integrity and a

residual load bearing capacity after the seismic event” (Eurocode 8 (1994)).

It is worth mentioning that, for structures showing the most critical damage (i.e. the 0.30g

structures), the local maximum values deviate from the average ones by a factor of about 1.5,

a) Configuration 2 b) Configuration 6

Direction YY

Direction XX

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Global Dam age Global Dam age

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.15g_L 0.15g_M 0.30g_M 0.30g_MSt 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.15g_L 0.15g_M 0.30g_M 0.30g_H

Intensity:

1.0 x Design

1.5 x Design

2.0 x Design

Page 407: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 383

which keeps approximately uniform regardless of the ductility class and of the intensity factor.

Such deviation does not appear exaggerated, particularly if one bears in mind the high degree

of redundancy of these statically indeterminate structures. Since the damage tends to locate

according to a beam sidesway mechanism, which, furthermore, does not develop at once by

virtue of the global structure post-yielding behaviour, a significant redistribution capacity can

be expected and, thus, a locally higher damage (e.g. 0.75, as herein obtained) is not likely to

strongly affect the structure safety.

The fact that structures of configuration 6, i.e., the irregular ones, typically show a trend for

similar (or even better) performance when compared to the regular ones (check for instance the

inter-storey drift and the average damage charts), suggests that the behaviour factor reduction

(0.8 as prescribed in EC8) reveals itself adequate to take into account the irregularity effects.

Moreover, although that reduction intended to account for the irregularity due to cut-off col-

umns, it ended up controlling also another irregular feature that is common to both configura-

tions 2 and 6. Actually, it was verified that the maximum damage mainly occurred in the

shorter span beams (as well as in the adjacent ground floor columns) of internal frames, contra-

rily to a more uniform distribution over the equal-span beams of the external frames. Further-

more, the damage patterns shown in Figure 7.12, suggest that the “irregularity” originated by

so different spans may be, at least, as important as the cut-off column irregularity.

Finally, a comment is due to the trial case 0.30g_Mst, for which the simplified static analysis

procedure was used in the design. This case is included mainly for comparison with the

0.30g_M structure of configuration 6, and, from the overall inspection of Figures 7.18 to 7.22,

a similar (or even better) performance is obtained for the 0.30g_Mst case. Similar (and some-

times lower) values of total drift, inter-storey drift and damage are obtained comparatively to

the 0.30g_M case, along with slightly higher global overstrength (see Figure 7.7), mostly aris-

ing from a significant overstrength at yielding (Table 7.6); these facts sustain the reliable fea-

tures of that simplified procedure.

7.4.6 Safety assessment by probabilities of failure. An exercise

In the preceding sections, the seismic performance of the structures studied herein was ana-

lysed in comparative terms by means of typical response variables. However, it is widely

accepted that, for a quantitative assessment of the structural safety recourse should be made to

Page 408: Seismic Assessment of Reinforced Concrete Frame Structures

384 Chapter 7

the probability of failure or, in more general terms, the probability of attaining a specific limit

state. Therefore, a first attempt was made to apply system reliability analysis to the structures

under study, in order to estimate probabilities of failure.

The complexity of the problems involved in system reliability analysis and of the correspond-

ing solution techniques would require a description and discussion at length, far beyond the

present work scope. No deep insight is envisaged into such issues, extensively addressed and

discussed by Pinto (1997), but the basic topics and assumptions are briefly presented and the

inherent difficulties are highlighted next. In this context, the basic aim is to have a measure of

safety against failure under seismic loading, mostly for comparative analysis between the dif-

ferent structures, rather than an absolute evaluation of the structural safety.

7.4.6.1 Methodology and assumptions

Local probability of failure

The computation of the global probability of failure for a given structure is commonly based

on the knowledge of local probabilities of failure in the structure critical zones where maxi-

mum damage is likely to develop (plastic hinges).

The calculation of hinge probability of failure requires a number of steps and assumptions

described next according to Figure 7.23. A main issue illustrated in that figure is the so-called

vulnerability function (Costa (1989), Duarte et al. (1990)) relating the seismic

action intensity (I) with the action effects (D), the damage index in the present case.

Figure 7.23 Local probability of failure at the hinge level

D ℑ I( )=

pa I( )

I

I∗

D∗

pr D( )

ps D( )

D

Page 409: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 385

The peak ground acceleration is selected as the action intensity, whose probabilistic quantifica-

tion is assumed described by the probability density function consistent with the site

seismicity.

The vulnerability function allows to map the probabilistic description of the action into the

space of action effects, viz the damage axis. Thus, the probability density function of damage

can be obtained, from which the probability of exceeding a given damage limit state

(say ) can be directly calculated as

(7.1)

where is the probability distribution of D.

Assuming the damage capacity described by the probability density function , the hinge

probability of failure is given by the following convolution integral (Campos Costa

(1993))

(7.2)

that can be seen as the result of the following reasoning:

• the damage capacity is less than or equal to D, the probability of which is given by

(7.3)

• the damage demand (action effect) falls in the interval , the corresponding

probability being

(7.4)

• the elementary probability of failure results from the intersection of the two above events

(7.5)

which, upon integration, leads to .

pa I( )

ps D( )

D∗

P D D∗>( ) 1 Fs D∗( )–=

Fs

pr D( )

Pf( )

Pf pr r( ) rd0

D

∫ ps D( ) Dd0

∫=

P r D≤( ) Fr D( ) pr r( ) rd0

D

∫= =

D D dD+[ , ]

P D s D dD+<≤( ) ps D( ) Dd=

dPf P r D≤( )P D s D dD+<≤( ) pr r( ) rd0

D

∫ ps D( ) Dd= =

Pf

Page 410: Seismic Assessment of Reinforced Concrete Frame Structures

386 Chapter 7

Note that Eq. (7.2) can be seen as the result of an integration in the action effect space (or

rather, axis) because the function is used in the outermost integral; however, the same

result could be obtained in the damage capacity space by the following alternative reasoning:

• the damage capacity falls in the interval , the probability of which is given by

(7.6)

• the damage demand (action effect) is greater than D and the corresponding probability is

(7.7)

• as before, the elementary probability of failure can be obtained and, consequently, the

total hinge probability of failure is given by

(7.8)

This alternative way of calculating is more convenient because it avoids the transformation

of into . Actually, for a given intensity such that , the following

equality holds

(7.9)

and can be transformed into

(7.10)

which means that, once defined the probability distributions of both the action intensity and of

the damage capacity, the value of becomes directly given by

(7.11)

Probabilistic quantification of the seismic intensity

For the seismic intensity quantification recourse was made to broad studies by Campos Costa

ps D( )

D D dD+[ , ]

P D r D dD+<≤( ) pr D( ) Dd=

P s D>( ) 1 Fs D( )– 1 ps s( ) sd0

D

∫–= =

dPf

Pf 1 Fs D( )–[ ]pr D( ) Dd0

∫=

Pf

pa I( ) ps D( ) I∗ I∗ ℑ 1– D∗( )=

P D D∗≤( ) P I I∗≤( )=

1 Fs D∗( )– 1 Fa ℑ 1– D∗( )( )–=

Pf

Pf 1 Fa ℑ 1– D( )( )–[ ]pr D( ) Dd0

∫=

Page 411: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 387

and Pinto (1997) and Pinto (1997) which attempt to characterize the European seismic hazard

scenarios, based on a large seismic database catalogue. The seismic hazard is categorized into

five different classes of increasing severity ranging from Very Low to High; for each seismic-

ity class, hazard estimates were obtained relating a series of the return periods (T) with the

expected peak ground accelerations and Weibull distributions were adopted to represent the

hazard curves.

For each class, the annual exceeding probability of peak ground acceleration (I) is given by

(7.12)

where the distribution parameters and are selected by imposing the hazard estimates for

return periods of 475 and 7000 years.

Two hazard curves were adopted herein, such that peak ground accelerations of 0.30g and

0.15g are obtained for the return period of 475 years (the reference one in the EC8). This corre-

sponded to the direct adoption of the High seismicity class (for approximately 0.30g) and of

scaled hazard from the Moderate High class to match the 0.15g acceleration at T=475 years.

The obtained curves are depicted in Figure 7.24 and the corresponding parameters are (αH =

0.376, βH = 0.00246) and (αM = 0.356, βM = 0.00093), respectively, for the high (0.30g) and

medium (0.15g) seismicities as considered herein.

Figure 7.24 High and medium seismicity hazard curves

P a I≥( ) 1T--- I

β---⎝ ⎠⎛ ⎞

α–exp= =

α β

10 100 1 103 1 1040

0.2

0.4

0.6

0.8

1

PGA (g)

T (years)

High Seismicity

Medium Seismicity

Page 412: Seismic Assessment of Reinforced Concrete Frame Structures

388 Chapter 7

Vulnerability functions and damage capacity

The vulnerability functions were fitted to the results of the non-linear analyses for the three

earthquake intensities, equal and above the design one. As schematically shown in Figure 7.25,

for a given plastic hinge, at each intensity four damage values were obtained (from the

response to the four earthquakes considered) and a curve of type

(7.13)

was fitted, where the parameters and were obtained by regression analysis; the corre-

sponding inverse becomes , with and , ready for

direct use in Eq. (7.11).

Figure 7.25 Vulnerability curve fitting to non-linear analysis results

Following an identical strategy as in previous reliability studies (e.g. Campos Costa (1993)),

for the damage capacity quantification a lognormal distribution of probability was adopted

(7.14)

where is the probability density of non-occurrence of failure for a damage value D. The

distribution parameters and can be related with the mean value and the coefficient of

variation cov by

(7.15)

D ℑ I( ) c1Ic2= =

c1 c2

I ℑ 1– D( ) b1Db2= = b1 1 c1⁄= b2 1 c2⁄=

I(g)

xxxx

xxxx

xxxx

D ℑ

I3I2I1

pr D( ) 12πδD

------------------

Dτ----⎝ ⎠⎛ ⎞

2ln

2δ2-----------------–exp=

pr D( )

τ δ D

δ 1 cov2+( )ln= τ D δ2

2-----⎝ ⎠⎛ ⎞exp⎝ ⎠

⎛ ⎞⁄=

Page 413: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 389

which, from the statistical analysis carried-out by Park et al. (1984) on a large set of experi-

mental results on reinforced concrete elements tested up to failure, can be taken approximately

as and .

System probability of failure

The system reliability, aiming at an estimate of the global probability of failure by combination

of local probabilities of failure, is one of the most complex and problematic issues in the struc-

tural safety evaluation. Notwithstanding the uncertainties related with both the seismic input

and the damage capacity characterization, the difficulty in establishing the combination of

local failure modes is a major obstacle due to its dependence on several aspects, viz the type of

loading (directly influencing the mechanisms of failure) and the correlation between action

effects and between damage capacities. Particularly, the latter aspects transform the evaluation

of the global probability of failure into a multi-variate problem formulated in a n-dimensional

space of random variables (where both action effects and resistances are included), whose ana-

lytical solution can hardly be obtained. Consequently, numerical approximations have to be

used, as for example the so called first and second order reliability bounds which consist of

estimates of lower and upper limits for the system probability of failure, rather than its “exact”

value.

In this line, structures are usually classified into series systems, or parallel systems or combina-

tions of both. The first type stands for systems where failure occurs when at least one of its

components fails, also designated by weakest-link systems (as suggested by the chain anal-

ogy). Instead, for parallel systems, failure is assumed to occur when all the components fail.

For series systems, the global probability of failure is limited by the Cornell bounds (Cor-

nell (1967), Madsen et al. (1986)):

(7.16)

where stands for the probability of failure of the ith component and n is the number of set

components. The lower bound states that the system fails when the most loaded component

fails, for full dependence between local failure events, while the upper bound, assuming com-

D 1= cov 0.5=

PF

max Pif( ) PF 1 1 Pi

f–( )i 1=

n

∏–≤ ≤

Pif

Page 414: Seismic Assessment of Reinforced Concrete Frame Structures

390 Chapter 7

plete independence of events, is obtained as the complementary of the system survival proba-

bility which equals the probability that all components survive

. Closer bounds can be defined for (Madsen et al. (1986), Pinto (1997)),

constituting the second order bounds, already including joint probability of events.

First order bounds for parallel systems can be written as , where the upper

bound identifies global failure with the failure of the less loaded component.

For practical purposes, statically determinate structures can be simulated as series systems, but

for the cases of major concern, i.e. statically indeterminate structures generally with high

degree of redundancy, the direct analogy with series or parallel systems is not apparent and,

most likely, not possible.

Indeed, for redundant structures, once the most loaded component fails, the redistribution

capacity allows the structure to “survive” further on, which means that, in an attempt for clas-

sifying it as a series system, the lower bound becomes overestimated: the structure

is “allowed” to have a lower value. Looking also at the upper bound, and bearing in mind

that it derives from the complementary probability of the intersection of survival events

assumed uncorrelated, the actual dependence between hinge events tends to reduce the value

of such bound. Thus, from this qualitative reasoning, it can be concluded that bounds of failure

probability for a series system constitute overestimates of bounds for a redundant structure.

On the other hand, the consideration of a redundant structure as a parallel system leads to

underestimated bounds of . Actually, to associate the system failure with the less loaded

component failure does not seem safe because of the different loading conditions that may trig-

ger the failure of distinct hinges (e.g., when the plastic hinges of structure upper storeys are

loaded to failure, typically the lower storey hinges may have already failed and endangered the

overall stability).

For these reasons, more sophisticated approaches can be adopted by means of parallel and

series systems combinations as proposed by Pinto (1997). Several mechanisms of failure are

previously identified under distinct load patterns assumed “independent” by virtue of imposed

vector orthogonality conditions. Such mechanisms are considered as parallel systems whose

failure probability is obtained from the local probabilities of failure of the relevant hinges (i.e.

only those activated for a specific mechanism). The global probability of failure is the outcome

PS 1 PF–=( )

PS 1 Pif–( )∏=( ) PF

0 PF min Pif( )≤ ≤

max Pif( )( )

PF

PF

Page 415: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 391

of a series association of those mechanisms, for which second order bounds are computed tak-

ing into account the correlation between mode failure events.

In the present work context, such an elaborated procedure is considered too much demanding

(in computational terms) for the envisaged purposes. Thus, a simple methodology was adopted

in which the design assumed mechanism is considered to control the failure mode (a beam

sidesway mechanism) and, despite the above mentioned overestimation trend, the Cornell

bounds were considered for the event set consisting of plastic hinge failure in all beams and in

the base end-zones of ground floor columns. Therefore, once the local probabilities of failure

are obtained by Eq. (7.11) for all the hinges considered, the bounds of global probability of

failure can be computed according to Eq. (7.16).

7.4.6.2 Comparative analysis

The above described methodology was applied for all trial cases under analysis and the

obtained upper and lower bounds of global annual probability of failure are plotted in the loga-

rithmic scale charts of Figure 7.26.

Figure 7.26 Bounds of annual probability of failure

PF

a) Configuration 2 b) Configuration 6

Direction YY

Direction XX

Design Case

1.00E-06

1.00E-05

1.00E-04

1.00E-03

1.00E-02

1.00E-01

1.00E+000.15g_L 0.15g_M 0.3g_M 0.3g_H

Limits:

Upper

Low er

Design Case

1.00E-06

1.00E-05

1.00E-04

1.00E-03

1.00E-02

1.00E-01

1.00E+000.15g_L 0.15g_M 0.3g_M 0.3g_Mst 0.3g_H

Limits:

Upper

Low er

Design Case

1.00E-06

1.00E-05

1.00E-04

1.00E-03

1.00E-02

1.00E-01

1.00E+000.15g_L 0.15g_M 0.3g_M 0.3g_H

Limits:

Upper

Low er

Design Case

1.00E-06

1.00E-05

1.00E-04

1.00E-03

1.00E-02

1.00E-01

1.00E+000.15g_L 0.15g_M 0.3g_M 0.3g_Mst 0.3g_H

Limits:

Upper

Low er

Page 416: Seismic Assessment of Reinforced Concrete Frame Structures

392 Chapter 7

As expected, high probabilities of failure are obtained, particularly in what concerns the upper

bound, with significant differences between the upper and lower bounds (in general over one

order of magnitude). For comparison purposes, it should be mentioned that values of annual

probability of failure around 2x10-4 are referred by Paulay and Priestley (1992) as appropriate

for the survival limit state of office buildings.

The structures designed for 0.15g exhibit some trend for lower probabilities of failure (10-4 to

10-3) than the 0.30g cases (10-3 to 10-2), particularly in the direction XX, which is coherent

with the less damage found in 0.15g structures (see Figures 7.21 and 7.22); additionally, it is

worth highlighting the reasonable uniformity of for these medium seismicity structures in

both directions XX and YY.

For 0.30g structures in the direction XX (where irregularities do exist due to both the cut-off

columns and the large span beams), the probability of failure tends to be higher than in the

direction YY where lateral loads are resisted by regular frames, which means that irregularities

contribute, as expected, for less safe solutions. Additionally, it is interesting to note that config-

uration 2, considered regular despite having adjacent beams with so different spans, shows

probabilities of failure higher than the assumed irregular structure, which sustains the ade-

quacy of the 80% reduction of the q-factor to account for irregularity in configuration 6. This

reduction, besides taking into account the cut-off column irregularity, also contributes to soften

the negative effects of the significant contrast of beam spans whose influence is quite apparent

in configuration 2.

The 0.30g_Mst and 0.30g_M cases (configuration 6) show quite similar probabilities of fail-

ure, which confirms the adequacy of the simplified static analysis procedure allowed in EC8.

Last but not the least, an interesting and important result is that, for a given design accelera-

tion, the ductility class does not seem to affect the structure reliability, which is fundamental

from the design code standpoint. Indeed, the EC8 key issue of allowing the designer to chose

between different ductility classes has to ensure that identically safe solutions are obtained,

and the present results seem to confirm the fulfilment of such requirement.

PF

Page 417: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 393

7.5 Concluding remarks

This section has dealt with the presentation and discussion of a numerical seismic assessment

study of some reinforced concrete frame structures designed according to EC8. The study,

included in the framework of the PREC8 activities, has focused on the non-linear seismic anal-

ysis of two building configurations (2 and 6) designed for several combinations of ductility

class and design acceleration (0.15g and 0.30g).

Besides an overview of the PREC8 programme in order to establish the present study context,

some insight was given into particular details of the structural configurations analysed herein:

the configuration 2, considered the regular one, and the configuration 6, where irregularities

were introduced by columns removed in the ground floor and by increased first storey height.

Structures were modelled with the flexibility element developed in the present work and non-

linear dynamic analyses were performed for each trial case under the action of four accelero-

grams fitting the EC8 spectrum with some increasing intensities.

Overall, the structures have shown an important reserve of strength which has been confirmed

by the seismic analysis results and by additional pushover static analyses. The available over-

strength at yielding (about 1.36) induces the 0.15g designed structures to behave in the pre-

yielding range, or just at incipient yielding, when loaded for the design seismic intensity. Glo-

bal overstrength factors as high as 2.1 can be found for earthquake action at twice the design

intensity, resulting from material overstrength and from global hardening.

Extensive cracking develops along structural members, leading the most significant stiffness

drop to take place at the design intensity. On the other hand, for the cases where significant ine-

lastic excursions occur (0.30g designed structures for ductility classes M and H), the local duc-

tility demands at the design intensity are lower than what could be expected, which results

from the available overstrength. Plastic hinging is reasonably spread all over the structure and,

for twice the design intensity, the beam sidesway mechanism shows up quite clearly. However,

the large contrast of span lengths of internal frames in one direction leads to some damage con-

centration in the shorter span central beams, which has been found as a source of irregularity

(actually not specifically foreseen and not taken into account in the assumed regular structure).

The total drift does not significantly differ between distinct ductility classes for the same

Page 418: Seismic Assessment of Reinforced Concrete Frame Structures

394 Chapter 7

design acceleration, although a slight trend can be observed for larger drifts with increasing

ductility class. Yet, low drifts are obtained (1.6%) for twice the design intensity, showing a

large margin to failure. The same applies for the interstorey drift, although the EC8 limit

observed in the design is somewhat exceeded by the obtained results (for the design intensity)

due to the extent of cracking development. However, no apparent second order effects are

likely to occur since the prescribed EC8 limit for the sensitivity coefficient is never exceeded.

In general, lower average damage is obtained when the ductility class is increased for the

same design acceleration, as a clear effect of the more stringent design provisions for ductility

enhancement. Typically, the 0.15g structures lead to low damage (about 0.1 for the design

intensity and less than 0.3 for twice that intensity), while in 0.30g cases the average damage

slightly exceeds 0.45 for twice the design intensity. These values are considered quite accepta-

ble, since they are close to the damage values obtained from similar numerical simulations of

the ELSA building structure for approximately identical seismic action levels and for which no

strong damage was detected in the experiment.

From a comparative study on probabilities of failure, structures designed for a given peak

ground acceleration have been found identically safe regardless of the ductility class they

belong to. This is an important issue since it appears to confirm the uniform risk assumption

underlying the designer “freedom” to trade between strength and ductility, as allowed in EC8

through the availability of three Ductility Classes. From this point on, the advantage of one

ductility class over the others has to be judged upon economical criteria, which constitutes a

major concern of a recent study by Pinto (1997).

Based on previous items (drift, damage and probabilities of failure) the following conclusions

are further highlighted:

• The behaviour factor reduction (80%) has shown quite good adequacy to account for the

irregularity source specifically considered in the design, since both the structural perform-

ance and the safety level of the irregular structures appeared as good as those of the regular

ones. In addition, it proved to be very important in reducing the negative effects of the non-

uniformity of beam spans which was present in both the regular and the irregular configura-

tions; possibly, this irregularity source should be considered for future inclusion in the irreg-

ularity criteria of EC8.

• The simplified static analysis method allowed in EC8 led to good results and to solutions as

Page 419: Seismic Assessment of Reinforced Concrete Frame Structures

SEISMIC ASSESSMENT OF RC FRAME STRUCTURES DESIGNED ACCORDING TO EC8 395

safe as those obtained by the reference multi-modal analysis method.

• Structures designed for high seismicity (0.30g), not only have shown higher demands (dam-

age included), but also led to larger probabilities of failure when compared with the medium

seismicity (0.15g) structures; this is related with the less explored non-linear behaviour in

the latter structures, for which the load combinations dominated by gravity loads are more

likely to control the design.

Overall, the analysed structures have shown quite good performance, well within the basic

requirements of EC8, and the main goals of the present study were reasonably achieved,

despite the improvement needs on specific topics (e.g. damage characterization and probability

of failure quantification).

Page 420: Seismic Assessment of Reinforced Concrete Frame Structures

396 Chapter 7

Page 421: Seismic Assessment of Reinforced Concrete Frame Structures

Chapter 8

FINAL REMARKS

8.1 Summary and conclusions

The present work was devoted to i) the development of a global element model, computation-

ally efficient for the analysis of RC structures under monotonic or cyclic loads inducing non-

linear behaviour throughout various response stages, ii) the model validation against results of

experimental tests on a full-scale structure and iii) its intensive application to the seismic

assessment of structures, particular those designed according to Eurocode 8 (EC8).

The innovative model

The model was required to describe the progressive modifications of stiffness due to non-line-

arity spread inside the member, such that one structural member could be modelled by only

one element. To this end, the flexibility formulation was chosen for the present development in

view of its adequacy to overcome the non-availability of kinematic shape functions to approx-

imate the element deformation field for different non-linear stages. Despite the use of this less

common formulation, the element model is easy to incorporate in typical finite element non-

linear analysis algorithms; thus, emphasis was mainly put on the element model development

and implementation, rather than well established non-linear schemes at the global structure

level.

Since the frame structural response to lateral actions is mainly of interest, the inelastic behav-

iour was essentially assumed at the element end zones, although cracking effects due to verti-

cal static loads were also taken into account in the span. The non-linear behaviour was

controlled in terms of moment-curvature by a modified Takeda-type model prescribed at the

element end sections; the non-linearity was assumed to spread along the element by recourse to

Page 422: Seismic Assessment of Reinforced Concrete Frame Structures

398 Chapter 8

cracking and yielding sections which are characterized by the cracking and yielding points,

respectively, in the trilinear skeleton curve of the model. Thus, at each load step, the element

was divided into distinct behaviour zones delimited by two types of sections, viz the fixed ones

(at member ends and at mid-span) and the moving sections (the cracking, yielding and null-

moment sections). Zones can be yielded (or plastic), cracked or uncracked; while the yielded

zones are allowed to develop only near the end sections, the cracked zones are considered

wherever the bending moment has exceeded the cracking threshold(s).

For a feasible and efficient (though approximate) control of cracked section zones, which are

likely to develop along considerable portions of the element, some modifications were intro-

duced in the adopted Takeda-type model such that a sudden transition from the uncracked to

the fully-cracked secant stiffness occurs once the cracking moment is exceeded. In doing so,

the element stiffness is progressively affected by the fully-cracked behaviour of cracked zones,

though no residual (permanent) deformations due to cracking are considered there for the

present stage of model development; this means that the hysteretic energy dissipation is not

properly considered in cracked zones.

In compliance with the general flexibility formulation, an internal iterative scheme was

required to accomplish the element state determination, driven by progressive elimination of

element displacement residuals and giving rise to the element end forces and tangent flexibility

matrix. However, by contrast with the general formulation, the displacement residuals are eval-

uated by integrating total deformations rather than residual ones since this was found more

suitable for the moving nature of most control sections. The control of plastic end zones

required particular care in order to minimize possible inaccuracies arising from the fact that

their behaviour is based only on the end and the yielding section behaviours; in addition, the

cracking development with the adopted uncracked/cracked transition was found to require par-

ticular detail for a successful advancement of the incremental-iterative scheme.

Implementation and auxiliary tools

The proposed element model was implemented in the general purpose computer code

CASTEM2000, an object-oriented program where both the well known Newton-Raphson

algorithm for non-linear system solution and the classical Newmark integration scheme were

available. For the particular nature of CASTEM2000, the basic code features were briefly

Page 423: Seismic Assessment of Reinforced Concrete Frame Structures

FINAL REMARKS 399

reviewed and the major interventions for the new element model implementation were pointed

out. The quite cumbersome, but unavoidable task of defining the moment-curvature trilinear

basic curve was pursued by means of a specifically developed algorithm for the most common

RC sections, viz the rectangular and T-shape ones. Designed to overcome the need of fibre dis-

cretization of the section, this algorithm was also implemented in CASTEM2000 as an auxil-

iary tool, where realistic behaviour models are considered for both steel and concrete, the latter

in confined or unconfined conditions; the section equilibrium is analysed for the curve turning

points, viz cracking, yielding and ultimate (or possibly other supplementary ones) under spe-

cific and pre-established point definition criteria.

The model validation and performance

Several experimentally tested cantilever beams having rectangular and T-shape cross-sections

were modelled by one flexibility element, for its validation at the single element level. Results

showed good numerical simulations of the experimental behaviour but some modelling limita-

tions were detected in relation to the hysteretic rules of the adopted section model, which, how-

ever, can be modified without interfering with the element formulation.

The element model validation at the global structure level was carried out through the simula-

tion of experimental tests on a four-storey reinforced concrete frame structure. The full-scale

specimen was designed according to Eurocode 8 as a high ductility structure and subjected to a

series of unidirectional seismic tests (pseudo-dynamically performed) and quasi-static cyclic

ones, thoroughly described in the present work.

Under the low level seismic test (corresponding to serviceability conditions in the pre-yielding

range), the bare structure performed quite well, although generalized cracking was found and

confirmed by the significant reduction of the fundamental frequency (from 1.78 Hz to 1.27

Hz). However, no yielding evidence was observed and the maximum inter-storey drift approx-

imately agreed with the EC8 limit. Proceeding to the high level seismic test (intending to

induce net inelastic excursions), cracking developed further, particularly in the beam-column

joint panels (with a bi-diagonal pattern) and in the beam-column interface section where major

cracks formed and rebar yielding took place; low energy dissipation was exhibited after the

response peaks, as apparent from the very “pinched” force-deformation diagrams, which was

related to bar-slippage inside the joints. However, rotations in the beam end zones were reason-

Page 424: Seismic Assessment of Reinforced Concrete Frame Structures

400 Chapter 8

ably well distributed, except for the top-storey where rather low rotation values were found.

Despite the referred crack formation, no significant permanent damage (e.g. concrete crushing

and spalling) was visible.

The seismic tests on the structure infilled with masonry panels showed a net increase of the ini-

tial frequency and some global strength enhancement resulting from the presence of infills.

Due to the high level test, the uniformly infilled structure suffered serious damage in the two

first storey masonry panels, responsible for the very “pinched” force-deformation diagrams

and the inherent low energy dissipation. The soft-storey infilled configuration (only partially

infilled in the storeys above the first) exhibited damage concentration in the first storey where

energy dissipation mainly took place; the large ductility demands in that storey induced signif-

icant strength degradation resulting from the observed concrete cover spalling in the first sto-

rey columns, which required the repair of damaged zones before proceeding with further

testing.

The quasi-static final cyclic tests (up to a top displacement ductility factor around 8) induced

progressively heavier damage, particularly in the two lower storeys where concrete cover

crushing and spalling was observed and followed by stirrup failure and buckling of rebars. The

highly damaged state of the structure, responsible for net degradation of global strength and for

large inter-storey drifts (around 7%), showed that a near-failure stage was actually reached at

the end of the test.

Several numerical analyses of the above referred structure were performed by recourse to the

proposed flexibility element model in order to reproduce the experimental behaviour through-

out the different testing stages. For a detailed insight of the critical zones behaviour, the chord

rotation at each element end section was selected as the key parameter to compute local ductil-

ity demands and to estimate damage; the latter was calculated according to the widely used

Park and Ang damage index, for which the quantification of the relevant parameters, viz the

ultimate chord rotation, the dissipated hysteretic energy and the corresponding weighting fac-

tor, was addressed in detail.

Since all the experimental tests consisted in imposing storey displacements quasi-statically and

in measuring the restoring forces, static simulations were possible by prescribing the same dis-

placements in the numerical model, the results being compared with the experimental ones.

Page 425: Seismic Assessment of Reinforced Concrete Frame Structures

FINAL REMARKS 401

For the bare frame seismic tests and, particularly, once a stabilized cracked behaviour was

reached, quite good simulations were obtained despite some lack of energy dissipation result-

ing from the model specific features in the cracking range. However, for net yielding develop-

ment during the high level test, numerical results agreed very well with experimental ones,

showing good simulation of dissipated energy, yet slightly over-evaluated possibly due to the

non-consideration of the pinching effect caused by anchorage slippage phenomena. Still for the

high level test, the ductility and damage distributions highlighted the expected strong column -

weak beam dissipation mechanisms; the maximum ductility demand (4.8) appeared reasonably

low in view of the behaviour factor (5) and the intensity factor (1.5), and the maximum damage

(0.4) complied with the observed state of the structure when compared with an approximate

damage scale proposed in the literature.

For the test simulations of the infilled structure configurations, the infill panels were modelled

by diagonal struts following a previous work, in which extensive background studies were per-

formed in order to develop a diagonal strut model, to make its validation and to derive the cor-

responding parameters. The numerical static analysis made herein did not provide good

simulations of the experimental behaviour, possibly as a result of the difficulties inherent in the

estimates of infill model parameters; indeed, better simulations were obtained after and where

extensive cracking affected the infill behaviour, because the response became mostly control-

led by the reinforced concrete frames.

Despite the difficulties in simulating the final cyclic test arising from the repair of damaged

zones, the global strength was reasonably described; however, these simulations mainly served

to highlight that the present hysteretic model rule for the evolution of unloading stiffness

induced excessive degradation when large ductilities are attained, thus leading to lower energy

dissipation than the experiment.

Dynamic analysis simulations of the bare structure seismic tests were also performed, for

which the structural frequencies showed good agreement with experimental measurements.

The dynamic response in the post-cracking range was very reasonably described by the model

upon inclusion of viscous damping forces (of Rayleigh type) to compensate for the lack of

model hysteretic dissipation in that behaviour range; the viscous damping factor (1.8%), exper-

imentally measured before the low level test, proved to be adequate while the response was

dominated by cracking. When the steel behaviour mostly controlled the response, as in the

Page 426: Seismic Assessment of Reinforced Concrete Frame Structures

402 Chapter 8

high level test, the viscous damping became useless because the model was able to incorporate

the hysteretic dissipation; the best simulations of the experimental results were obtained with-

out viscous damping. However, it was found that, should viscous damping be included, its con-

tribution to the energy dissipation can be very significant, even for low damping factors.

The proposed element model performance was compared with that from analyses by tradi-

tional fixed length plastic hinge modelling. The comparison was restricted to the bare frame

seismic tests, having shown that, for static simulations, the traditional way of modelling leads

to results less good than the proposed element in the pre-yielding range, while both modelling

strategies give similar results for post-yielding analyses. In the dynamic context, the flexibility

element model led to better approximation of the experimental response as a result of the pro-

gressive adaptation of member stiffness, indeed the main advantage over the traditional fixed

hinge modelling. The dynamic behaviour was better simulated because the structural stiffness,

and consequently the vibration characteristics, were closely described and updated. It should

be emphasized that good simulations were obtained in the pre-yielding range by adopting the

fully-cracked section stiffness progressively introduced in the element along with viscous

damping forces, while in the post-yielding range the behaviour was better described by remov-

ing any viscous dissipation. Globally, the seismic behaviour was adequately simulated by the

proposed flexibility element modelling, though some issues are recognized to require further

development.

Analyses for Eurocode 8 validation

Aiming at the numerical seismic assessment of reinforced concrete frame structures designed

according to EC8, a number of non-linear seismic analyses of two building configurations

were carried out as part of a european-wide pre-normative research programme. Building

structures were designed for several combinations of ductility class and design accelerations

(0.15g and 0.30g, corresponding to medium and high seismicity zones, respectively); one con-

figuration was considered regular, while the other included irregularities caused by columns

removed in the ground floor and by increased first storey height. Each trial case was modelled

by recourse to the proposed flexibility element model and the non-linear analyses were carried

out under the action of four accelerograms fitting the EC8 spectrum and scaled by three

increasing intensities, at and above the design one (additional analyses at half the design inten-

sity were also performed for the high ductility class structures).

Page 427: Seismic Assessment of Reinforced Concrete Frame Structures

FINAL REMARKS 403

Significant overstrength was found for all structures resulting from the strength reserve of con-

stituent materials (where the partial safety factors for material properties are included), some

design requirements (minimum reinforcement, capacity design and detailing), the gravity load

influence and the post-yielding hardening at the material and the global structure levels. Over-

strength factors of about 1.36 were obtained at yielding, such that the 0.15g designed trial

cases exhibited pre-yielding behaviour, or just imminent yielding, under the design seismic

intensity; however, overstrength increases further with the seismic intensity, mainly resulting

from global hardening, and overstrength factors above 2.0 could be found for twice the design

intensity.

At the design intensity, extensive cracking was found in structural members leading to very

significant stiffness reductions, and, due to overstrength, the local ductility demands were

lower than expected in the cases where the most significant inelastic excursions occurred (i.e.

in the 0.30g designed structures). Despite the fact that a reasonable spread of plastic hinging

could be found throughout all the structures, some damage concentration appeared in the

shorter span beams of frames with contrasting span lengths; this was found to be a source of

irregularity not duly foreseen and not considered in the design of the assumed regular struc-

ture.

Both the total and the inter-storey drifts were not strongly affected by changing the ductility

class for the same design acceleration, though a slight trend was found showing larger drifts

when the ductility class is increased. However, even for twice the design intensity, low drifts

were obtained (maximum values of 1.6% and 2.3%, respectively for the total and the inter-sto-

rey drifts), at least three times below the values reached at failure in the experimentally tested

structure above referred; this fact indicated that a large margin to failure was available.

By contrast, a clear trend showed that lower average damage was obtained when the ductility

class was increased for a given design acceleration, as a consequence of the more stringent

design provisions to enhance ductility. However, the ductility class was found not to affect the

safety level of structures designed for the same peak ground acceleration, since not very differ-

ent failure probability bounds were obtained for distinct ductility classes. From the EC8 stand-

point this is a very important finding, because it contributes to ensure that, by trading between

strength and ductility, the designer obtains identically safe solutions.

Page 428: Seismic Assessment of Reinforced Concrete Frame Structures

404 Chapter 8

Globally, the analysed structures showed quite good performance within the EC8 require-

ments. The behaviour factor reduction to account for irregularities proved to be adequate and,

additionally, very important for reducing the negative effects induced by non-uniformity of

beam-spans. Good results and identically safe solutions were provided by the simplified static

analysis method allowed in EC8 when compared to those obtained by the reference multi-

modal analysis method. The less explored non-linear behaviour of medium seismicity (0.15g)

structures, for which the design is more likely to be controlled by load combinations domi-

nated by gravity forces, led to lower demands (including damage) and lower probabilities of

failure than structures designed for high seismicity (0.30g).

8.2 Future developments

In line with the work developed in the present thesis, a number of topics are deemed important

for further research regarding the improvement of modelling strategies and the refinement of

seismic assessment studies. Concerning the modelling issues, the following are highlighted:

• A softening branch in the post-yielding range should be included in the moment-curvature

section model, in order to allow a more realistic description of the near-failure stage; this is

of particular interest in columns due to the axial force effects. To this end, the proposed

algorithm to obtain the trilinear envelope curve could be extended for a better definition of

the post-yielding branch.

• The possibility of using an intermediate stiffness between the uncracked and the fully-

cracked ones in the moment-curvature relationship should be investigated in order to

account for the tension stiffening effect in the element cracked zones. Particularly, in the

presence of axial force, deviations between the actual post-cracking and the fully-cracked

stiffness may become significant, which should be further addressed and, possibly, compen-

sated by means of an adequate intermediate stiffness estimated on the basis of the actual

characteristics of the whole element.

• It has been clearly pointed out that both the unloading stiffness degradation and the pinch-

ing rules should be improved. A broad and systematic study of the numerical simulation of

the experimental response of single reinforced concrete members or sub-assemblages would

be extremely useful. Ideally, a reduced number of the most recent phenomenological mod-

els should be selected to provide a variety of hysteretic rules in order to use them in the

numerical simulations. On the other hand, a considerable amount of results of reinforced

Page 429: Seismic Assessment of Reinforced Concrete Frame Structures

FINAL REMARKS 405

concrete experimental activity is nowadays worldwide available which should be grouped

according to some basic variables such as, the section type, the steel and concrete grades,

the test type, the maximum deformation (or ductility) reached, the shear-span ratio, the

proneness to pull-out effects, etc. Such classification would contribute to clarify and to

render more consistent the fitting of hysteretic rules to the experimental evidence and the

definition of empirical relations between hysteretic parameters and the member characteris-

tics.

• In view of the difficulties in correctly modelling the hysteretic dissipation in cracked zones

continuously developing, and bearing in mind that viscous damping proved to be adequate

in the pre-yielding range when considered simultaneously with the progressive inclusion of

the fully-cracked stiffness in the member, a different and perhaps more consistent or general

way of including the viscous effect should be investigated. A possible solution could rely on

the adoption of a global damping matrix resulting from the assemblage of element damping

matrices, which in turn could be proportional to the element stiffness matrices where the

effects of cracked zone development would come directly included.

• A more tight control of yielded zones should be considered in order to closely follow the

actual flexibility and curvature distributions in such zones. For this purpose, recourse could

be made to the progressive activation of some fixed sections in the yielded zones to be fully

controlled by the same section model of the corresponding end section.

• The influence of inclined shear cracking in the yielded zones should be investigated, partic-

ularly concerning possible increase of inelasticity spread as suggested by the usual strut-

and-tie modelling for shear analysis of reinforced concrete members. The consequences of

further extending the yielded zones to account for shear and how much this extension

should be, deserve particular attention.

• The suitability of the flexibility formulation to handle associations in series of several ele-

ments should be explored to complement the proposed element with others devoted to the

simulation of rebar slippage and non-linear shear effects; the former could be taken into

account by fixed point hinges at member ends, while the latter could be described by an ele-

ment with two variable length extreme zones (where the non-linear shear effects develop)

connected by a rigid element.

Concerning the seismic assessment of reinforced concrete structures, the following topics are

emphasized:

Page 430: Seismic Assessment of Reinforced Concrete Frame Structures

406 Chapter 8

• Studies for a better characterization of damage should be undertaken; new proposals for

damage index computation at the critical zone level are needed and, despite the energy-

based proposals appear adequate, research is still lacking on how to split an element damage

index into its critical zones.

• The safety assessment by failure probability should be improved, particularly in what con-

cerns the probabilistic characterization of the damage capacity and the extension of vulner-

ability functions to higher levels of seismic intensity. The system reliability also requires a

deeper insight in order to obtain closer and more reliable bounds of the failure probability.

In addition, the safety assessment study should be extended to all structures considered in

the PREC8 programme, in order to check whether the present work conclusions also apply

to the other trial cases.

• Further research activity in line with the PREC8 programme should be undertaken to find

out the effects of irregularities in a systematic and objective way. To this end, not only the

safety assessment would be required, but also the quantification of structural overstrength

and effective behaviour factors would be helpful in assessing the EC8 rules for ductile seis-

mic design.

Page 431: Seismic Assessment of Reinforced Concrete Frame Structures

REFERENCES

A

Anagnostopoulos, S. A. (1972) - Nonlinear Dynamic Response and Ductility Requirements ofBuilding Structures Subjected to Earthquakes - Research Report No R72-54, Depart-ment of Civil Engineering, Massachusetts Institute of Technology, Cambridge, Ma.

Anthoine, A. (1997) - POP Sicilia, Progetto E, Modulo E10-c - Final Report, ISIS, JRC, Ispra,Italy (in italian).

Arêde, A. and Pinto, A. V. (1996) - Reinforced Concrete Global Section Modelling: Definitionof Skeleton Curves - Special Publication No.I.96.36, ISIS, EC, JRC, Ispra (VA), Italy.

Arêde, A., Campos Costa, A. and Pinto, A. V. (1996) - Non-linear Seismic Response of Build-ing Structures Designed in Accordance with EC2 and EC8 (Configurations 2 and 6) -Report EUR No.16356 EN, STI, CEC, JRC, Ispra (VA), Italy.

Arêde, A. and Pinto, A. V. (1997) - A New Flexibility-Based Global Element Model for RCFrame Analysis - Special Publication. ISIS, EC, JRC, Ispra (VA), Italy (in prepara-tion).

Argyris, J. H.; Vaz, L. E. and William, K. J. (1978) - Improved Solution Methods for InelasticRate Problems - Computational Methods in Applied Mechanics and Enginnering, Vol.16, pp. 231-277.

Arzoumanidis, S. G. and Meyer, C. (1981) - Modelling of Reinforced Concrete Beams Sub-jected to Cyclic Loads - Report NSF. CEE81-21359-1. Dep. of Civil Engineering andEngineering Mechanics, Columbia University, New York, N.Y.

B

Banon, H. (1980) - Prediction of Seismic Damage in Reinforced Concrete Frames - Publica-tion No. R80-16. Dep. of Civil Engineering, Massachusetts Institute of Technology,Cambridge, Massachusetts.

Banon, H. and Veneziano, D. (1982) - Seismic Safety of Reinforced Concrete Members and

Page 432: Seismic Assessment of Reinforced Concrete Frame Structures

408 REFERENCES

Structures - Earthquake Engineering and Structural Dynamics. Vol. 10.

Brancaleoni, F.; Ciampi, V. and Di Antonio, R. (1983) - Rate-Type Models for Non-LinearHysteretic Structural Behaviour - EUROMECH Coll., Palermo, Italy

Buchet, Ph. and Pegon, P. (1994) - PSD Testing with Sustructuring: Implementation and Use -Special Publication No.I.94.65. STI, CEC, JRC, Ispra (VA), Italy.

C

Campos Costa, A. (1993) - A Acção dos Sismos e o Comportamento das Estruturas - Disser-tação para Doutoramento em Engenharia Civil, Faculdade de Engenharia da Universi-dade do Porto, Porto (in Portuguese).

Campos Costa, A. and Pinto, A. V. (1997) - European Seismic Hazard Scenarios. AnApproach to the Definition of Input Motions for Testing and Reliability Assessment ofCivil Engineering Structures - ISIS, JRC, Ispra (VA), Italy (in preparation)

Calvi, G. M. and Pinto, P. E. (1996) - Experimental and Numerical Investigations on the Seis-mic Response of Bridges and Recommendations for Code Provisions - PREC8/ECOEST Report, LNEC, Lisboa.

Carvalho, E. C. (1991) - Cooperative Research Program on the Seismic Response of Rein-forced Concrete Structures - Contract 3919-90-02 ED ISP P, Final Report, LNEC, Lis-boa.

Carvalho, E. C. (1992) - Cooperative Research Program on the Seismic Response of Rein-forced Concrete Structures - Contract 4504-91-10 ED ISP P, Interim Report, LNEC,Lisboa.

Carvalho, E. C. (1993) - Cooperative Research Program on the Seismic Response of Rein-forced Concrete Structures - Contract 4504-91-10 ED ISP P, Final Report, LNEC, Lis-boa.

Carvalho, E. C., Coelho, E. and Fardis, M. (1996) - Assessment of EC8 provisions for rein-forced concrete frames - Proc. of the 11th World Conference on Earthquake Engineer-ing, Acapulco, Mexico.

CEA (1990) - CASTEM 2000, Guide d’utilisation - CEA, Saclay, France

CEB-FIP Model Code 1990, First Draft, Comité Euro-International du Béton, Bulletind´Information No 195 and 196, Lausanne.

Ciampi, V. and Carlesimo, L. (1986) - A Non-linear Beam Element for Seismic Analysis ofStructures - Proc. of the 8th European Conference on Earthquake Engineering, Lisboa,Portugal.

Clough, R. W.; Benuska, K. L. and Wilson, E. L. (1965) - Inelastic Earthquake Response of

Page 433: Seismic Assessment of Reinforced Concrete Frame Structures

REFERENCES 409

Tall Buildings - Proc. of the 3rd World Conference on Earthquake Engineering, NewZealand, Vol. 11.

Clough, R. W. and Johnston, S. B. (1966) - Effect of Stiffness Degradation on EarthquakeDuctility Requirements - Proc. of the Japan Earthquake Engineering Symposium,Tokio.

Clough, R. W. and Penzien, J. (1975) - Dynamics of Structures - McGraw-Hill Book Com-pany, Singapore.

Coelho, E. (1992) - Comportamento Sísmico de Estruturas em Pórtico de Betão Armado -Dissertação para Doutoramento em Engenharia Civil, Instituto Superior Técnico, Lis-boa, (in Portuguese).

Combescure, D. and Pegon, P. (1994) - Alpha-Operator Splitting Time Integration Techniquefor Pseudo-Dynamic Tests - Error Propagation Analysis - Special PublicationNo.I.94.65. STI, CEC, JRC, Ispra (VA), Italy.

Combescure, D. (1996) - Modélisation du comportement sous chargement sismique des struc-tures de bâtiment comportant des murs de remplissage en maçonerie - Ph.D. Thesis,École Centrale de Paris, Paris, (in French).

Cornell, C. A. (1967) - Bounds on the Reliability of Structural Systems - Journal of the Struc-tural Division, ASCE, Vol. 93, ST12, pp 171-200.

Costa, A.G. and Campos Costa, A. (1987) - Modelo Histerético das Relações Forças-Desloca-mentos Adequadas à Análise Sísmica de Estruturas - Relatório 269/87-NDA DE,LNEC, Lisboa (in Portuguese).

Costa, A.G. (1989) - Análise Sísmica de Estruturas Irregulares - Dissertação para Douto-ramento em Engenharia Civil, Faculdade de Engenharia da Universidade do Porto,Porto (in Portuguese).

Criesfield, M. A. (1982) - Accelerated Solution Techniques and Concrete Cracking - Compu-tational Methods in Applied Mechanics and Enginnering, Vol. 33, pp. 585-607.

D

Dahlblom, O. and Ottosen, S.N. (1991) - Smeared Crack Analysis using Generalized Ficti-tious Crack Model - Journal of Engineering Mechanics, ASCE Vol. 116, No. 1.

Delgado, R. M. (1984) - O Método dos Elementos Finitos na Análise Dinâmica de Barragens,Incluindo A Interacção Sólido-Líquido - Dissertação para Doutoramento em Engen-haria Civil, Faculdade de Engenharia da Universidade do Porto, Porto (in Portuguese).

Dias da Silva, V. (1995) - Mecânica e Resistência dos Materiais - Ediliber Eds., Coimbra, Por-tugal (in portuguese).

Page 434: Seismic Assessment of Reinforced Concrete Frame Structures

410 REFERENCES

Donea, J.; Jones, P.M.; Magonette, G.E. and Verzeletti, G. (1990) - The Pseudo-Dynamic TestMethod for Earthquake Engineering: An Overview - EUR 12846 EN. STI, CEC, JRC,Ispra (VA), Italy.

Donea, J.; Magonette, G. E.; Negro, P.; Pinto, A. V. and Verzeletti, G. (1996) - Pseudo-Dynamic Capabilities of the ELSA Laboratory for Earthquake Testing of Large Struc-tures - Earthquake Spectra, Vol. 12, No.1, pp. 163-180.

Duarte, R.; Oliveira, C. S.; Campos Costa, A. and Costa, A. G. (1990) - A Non-Linear Modelfor Seismic Analysis, Design and Safety Assessment of Reinforced Concrete Buildings- Earthquake Damage Evaluation & Vulnerability Analysis of Building Structures, A.Koridze, Omega, England.

E

Elnashai, A.S. and Papazoglou, A.J. (1995) - Vertical Earthquake Ground Motion. EvidenceEffects and Simplified Analysis Procedures - Research Report ESEE-95/6, ImperialCollege, London.

Elnashai, A.S. and Papazoglou, A.J. (1997) - Procedure and Spectra for Analysis of RC Struc-tures Subjected to Strong Vertical Earthquake Loads - Journal of Earthquake Engi-neering, Vol. 1, No 1, pp. 121-155.

ENV 10080 (1991) - Steel for the Reinforcement of Concrete - Weldable Ribbed ReinforcingStell B500 - Technical Delivery Conditions for Bars, Coils and Welded Fabric - ECISS/TC19/SC1.

Eurocode No 2 (1991) - Design of Concrete Structures, Part 1 - General Rules and Rules forBuildings, pr ENV 1992-1-1 - CEN/TC250/SC2.

Eurocode No 3 (1992) - Design of Steel Structures, Part 1-1 - General Rules and Rules forBuildings, pr ENV 1993-1-1 - CEN/TC250/SC3.

Eurocode No 4 (1992) - Design of Composite Steel and Concrete Structures, Part 1-1 - Gen-eral Rules and Rules for Buildings, pr ENV 1994-1-1 - CEN/TC250/SC4.

Eurocode No 8 (1994) - Design Provisions for Earthquake Resistant Structures, Part 1-1, 1-2,1-3, pr ENV 1998-1-1, 1-2, 1-3 - CEN/TC250/SC8.

F

Fardis, M.N. (1991) - Member-Type Models for the Non-Linear Seismic Response Analysisof Reinforced Concrete Structures - Experimental and Numerical Methods in Earth-quake Engineering. J. Donea & P.M. Jones eds. CEC, JRC, Ispra (VA), Italy.

Page 435: Seismic Assessment of Reinforced Concrete Frame Structures

REFERENCES 411

Fardis, M.N.; Economu, S.N.; Antoniou, A.N.; Komodromos, P.J. and Sfakianakis, M.G.(1993) - Damage Measures and Failure Criteria - Part I, Contribuition of University ofPatras, Final Report of Cooperative Research on the Seismic Response of ReinforcedConcrete Structures - 2nd Phase, LNEC, Lisboa.

Fardis, M. (1994) - PREC8, Prenormative research in support of Eurocode 8. Topic 1 - Rein-forced concrete frames and walls. Design of the trial cases - University of Patras. Pat-ras. Greece.

Fardis, M. (1995) - Current Trends in Earthquake Resistant Analysis and Design of Rein-forced Concrete Structures - Proc. of the SECED Conference - European SeismicDesign Practice, Chester, United Kingdom.

Fardis, M. and Panagiotakos, T. B. (1997) - Seismic Design and Response of Bare andMasonry-Infilled Reinforced Concrete Buildings. Part I: Bare Structures - Journal ofEarthquake Engineering, Vol. 1, No 1, pp. 219-256.

Faria, R. (1994) - Avaliação do Comportamento Sísmico de Barragens de Betão Através deum Modelo de Dano Contínuo - Dissertação para Doutoramento em Engenharia Civil,Faculdade de Engenharia da Universidade do Porto, Porto (in Portuguese).

Feenstra, P. H. (1993) - Computational Aspects of Biaxial Stress in Plain and Reinforced Con-crete - Ph.D. Thesis, Technical University of Delft, Delft.

Figueiras, J. A. (1983) - Ultimate Load Analysis of Anisotropic and Reinforced ConcretePlates and Shells - Ph.D. Thesis, University of Swansea, Swansea.

Filippou, F. C.; Popov, E. P. and Bertero V. V. (1983) - Effects of Bond Deterioration on Hys-teretic Behaviour of Reinforced Concrete Joints - Report No. UCB/EERC 83-19, Earth-quake Engineering Research Center, University of California, Berkeley.

Filippou, F. C. and Issa, A. (1988) - Non-linear Analysis of Reinforced Concrete Framesunder Cyclic Load Reversals - Report No. UCB/EERC 88-12, Earthquake EngineeringResearch Center, University of California, Berkeley.

Filippou, F. C.; D’Ambrisi, A. and Issa, A. (1992) - Non-linear Static and Dynamic Analysisof Reinforced Concrete Subassemblages - Report EERC 92-08, Earthquake Engineer-ing Research Center, University of California, Berkeley.

G

Giberson, M. (1967) - The Response of Nonlinear Multi-Story Structures Subjected to Earth-quake Excitations - Earthquake Engineering Research Laboratory, Pasadena.

Guedes, J.; Pegon, P. and Pinto; A. V. (1994) - A Fibre/Timoshenko Beam Element inCASTEM 2000 - Special Publication No.I.94.31. STI, CEC, JRC, Ispra (VA), Italy.

Gutierrez, E., Verzeletti, G. and Magonette, G. E. (1993) - Experimental Studies of Loading

Page 436: Seismic Assessment of Reinforced Concrete Frame Structures

412 REFERENCES

Rate Effects on Reinforced Concrete Columns - Journal of Structural Engineering,ASCE, Vol. 119, No. 5.

H

Hellesland, J. and Scordelis, A. (1981) - Analysis of RC Bridge Columns Under ImposedDeformations - IABSE, Delphi, pp. 545-559.

Hilber, H. M.; Hughes, T. J. R. and Taylor, R. L. (1977) - Improved Numerical Dissipation forTime Integration Algorithms in Structural Dynamics - Earthquake Engineering andStructural Dynamics, Vol.5, pp 283-292.

Hillerborg, A., Modeer, M. and Petersson, P. E. (1976) - Analysis of Crack Formation andGrowth in Concrete by Means of Fracture Mechanics and Finite Elements - Cementand Concrete Research, Vol. 6, pp. 773-782.

Hughes, T. J. R.; Pister, K. S. and Taylor, R. L. (1979) - Implicit-explicit finite elements innonlinear transient analysis - Computer Methods in Applied Mechanical Engineering,pp.159-182.

Hughes, T. J. R. (1987) - The Finite Element Method - Linear Static and Dynamic Finite Ele-ment Analysis - Prentice-Hall, Inc., Englewood Cliffs, New Jersey 07632.

I

Izzuddin, B. A. (1991) - Non-linear Dynamic Analysis of Framed Structures - Ph.D. Thesis,Imperial College, London.

Izzuddin, B. A.; Karayannis, C. G. and Elnashai, A. S. (1994) - Advanced Non-linear Formu-lation for Reinforced Concrete Beam-Columns - Journal of Structural Engineering,ASCE, 120, No. 10, pp. 2913-2934.

J

Jennings, P. (1963) - Response of Simple Yielding Structures to Earthquake Excitation - Ph.D.Thesis, Pasadena, California.

K

Kaba, S. and Mahin, S. A. (1984) - Refined Modelling of Reinforced Concrete Columns for

Page 437: Seismic Assessment of Reinforced Concrete Frame Structures

REFERENCES 413

Seismic Analysis - Report EERC 84-03, Earthquake Engineering Research Center,University of California, Berkeley.

Kakaliagos, A. (1994) - Pseudo-Dynamic Testing of a Full-Scale Three-Storey One-Bay SteelMoment-Resisting Frame. Experimental and Analytical Results - Report EURNo.15605 EN. STI, EC, JRC, Ispra (VA), Italy.

Karayannis, C. G.; Izzuddin, B. A. and Elnashai, A. S. (1994) - Application of Adaptive Anal-ysis to Reinforced Concrete Frames - Journal of Structural Engineering, ASCE, Vol.120, No. 10, pp. 2935-2957.

Kunnath, S. K.; Reinhorn, A. M. and Park, Y. J. (1990) - Analytical modelling of inelasticseismic response of R/C structures - Journal of Structural Engineering, ASCE Vol.116, No. 4.

Kunnath, S. K.; Reinhorn, A. M. and Lobo, R. F. (1992) - IDARC Version 3.0: A Program forthe Damage Inelastic Analysis of Reinforced Concrete Structures - Technical reportNCEER-92-0022. State University of New York at Buffalo-Buffalo, New York 14260and University of Central Florida, Orlando, Florida 32816-0450.

L

Litton, R. W. (1975) - A Contribution to the Analysis of Concrete Structures under CyclicLoading - Ph.D. Thesis, Department of Civil Engineering, University of California,Berkeley.

M

Madsen, H. O.; Krenk, S. and Lind, N. C. (1986) - Methods of Structural Safety - Prentice-Hall, Inc., Englewood Cliffs, New Jersey 07632.

Magonette, G. E. (1991) - Digital Control of Pseudo-Dynamic Tests - Experimental andNumerical Methods in Earthquake Engineering. J. Donea & P.M. Jones eds. CEC,JRC, Ispra (VA), Italy.

Mahasuverachai, M. (1982) - Inelastic Analysis of Piping and Tubular Structures - ReportEERC 82-27, Earthquake Engineering Research Center, University of California, Ber-keley.

Mari, A. and Scordelis, A. (1984) - Nonlinear Geometric Material and Time Dependent Anal-ysis of Three Dimensional Reinforced and Prestressed Concrete Frames - SESMReport 82-12, Department of Civil Engineering, University of California, Berkeley.

Marques, J. M. M. C. (1986) - Nonlinear Finite Element Solutions with Quasi- and Secant-Newton Methods - Proc. of International Conf. on Numerical Methods for Nonlinear

Page 438: Seismic Assessment of Reinforced Concrete Frame Structures

414 REFERENCES

Problems, Dubrovnik, Youguslavia, C. Taylor, D. R. J. Owen, E. Hinton & F. B. Dan-jaric Eds., Pineridge Press.

Matheis, H. and Strang, G. (1979) - The Solution of Nonlinear Finite Element Equations -International Journal of Numerical Methods in Engineering, Vol. 14, pp. 1613-1626.

Menegotto, M. and Pinto, P. (1973) - Methods of Analysis for Cyclic Loaded Reinforced Con-crete Plane Frames Including Changes in Geometry and Non-elastic Behaviour of Ele-ments under Combined Normal Force and Bending - IABSE Symposium on resistanceand ultimate deformability of structures acted on by well-defined loads, Final report,Lisboa.

Menegotto, M. and Pinto, P. (1977) - Slender Reinforced Concrete Compressed Members inBiaxial Bending - Journal of Structural Engineering, ASCE, 103 (ST3), pp. 587-605.

Meyer, C., Roufaiel, M. S. L. and Arzoumanidis, S. G. (1983) - Analysis of Damaged Con-crete Frames for Cyclic Loads - Earthquake Engineering and Structural Dynamics,Vol. 11, pp 207-228.

Meyer, C., Krätzig, W. B., Stangenberg, F. and Meskouris, K. (1988) - Damage Prediction inReinforced Concrete Frames Under Seismic Actions - European Earthquake Engineer-ing, Vol. 2, No. 3, pp 9-15.

Monti, G.; Spacone, E. and Filippou, F. C. (1993) - Model for Anchored Reinforcing BarsUnder Seismic Excitations - Report EERC 93-08, Earthquake Engineering ResearchCenter, University of California, Berkeley.

N

Nakashima, M.; Kaminosono, T.; Ishida, M. and Ando, K. (1990) - Integration Techniques forSubstructure Pseudo-Dynamic Test - Proc. of Conference on Earthquake Engineering,Palm Springs CA, USA.

Negro, P.; Verzeletti, G.; Magonette, G. E. and Pinto, A. V. (1994) - Tests on a Four-StoreyFull-Scale RC Frame Designed According to Eurocodes 8 and 2: Preliminary Report -Report EUR No.15879 EN. STI, CEC, JRC, Ispra (VA), Italy.

Negro, P.; Anthoine, A.; Combescure, D.; Magonette, G. E.; Molina, J.; Pegon, P. andVerzeletti, G. (1995) - Tests on the Four-Storey Full-Scale Reinforced Concrete Framewith Masonry Infills: Preliminary Report - Special Publication No.I.54.54. STI, CEC,JRC, Ispra (VA), Italy

O

Otani, S. (1974) - Inelastic Analysis of RC Frame Structures - Journal of Structural Engineer-

Page 439: Seismic Assessment of Reinforced Concrete Frame Structures

REFERENCES 415

ing, ASCE, Vol. 100, ST7.

Owen, D. R. J. and Hinton, E. (1980) - Finite Elements in Plasticity - Pineridge Press, Lim-ited, Swansea, U.K.

Ozdemir, H. (1981) - Nonlinear Transient Dynamic Analysis of Yielding Structures - Ph.D.Thesis, Department of Civil Engineering, University of California, Berkeley.

P

Papazoglou, A. J. and Elnashai, A. S. (1996) - Analytical and Field Evidence of the DamagingEffect of Vertical Earthquake Ground Motion - Earthquake Engineering and StructuralDynamics, Vol.25, pp 1109-1137.

Park, R. and Paulay, T. (1975) - Reinforced Concrete Structures - John Wiley & Sons, Inc.,New York.

Park, Y. J.; Ang, A. H-S. and Wen, Y. K. (1984) - Seismic Damage Analysis and Damage-lim-iting Design of RC Buildings - Technical report No. UILU-ENG-84-2007, Universityof Illinois at Urbana-Champaig, Urbana, Illinois.

Park, Y. J.; Ang, A. H-S. and Wen, Y. K. (1987) - Damage-Limiting Aseismic Design ofBuildings - Earthquake Spectra, Vol. 3, No. 1.

Park, Y. J.; Reinhorn, A. M. and Kunnath, S. K. (1987) - IDARC: Inelastic Damage Analysisof Reinforced Concrete Frame-Shear-Wall Structures - Technical report NCEER-87-0008, State University of New York at Buffalo-Buffalo, New York.

Park, R., Priestley, M. J. N. and Gill, W. D. (1990) - Ductility of Square Confined ConcreteColumns - Journal of Structural Engineering, ASCE, Vol. 108, No. 12.

Paulay, T. and Priestley, M. J. N. (1992) - Seismic Design of Reinforced Concrete andMasonry Buildings - John Wiley & Sons, Inc., New York.

Pegon, P. (1993) - A Timoshenko Simple Beam Element in CASTEM 2000 - Technical NoteNo. I.93.05, STI, CEC, JRC, Ispra (VA), Italy.

Pegon, P. (1996) - PSD Testing with Substructuring: the Case of Asynchronous Motion - Spe-cial Publication No.I.96.42, ISIS, EC, JRC, Ispra (VA), Italy.

Pegon, P. and Anthoine, A. (1994) - Numerical Strategies for Solving Continuum DamageProblems Involving Softening: Application to the Homogenization of Masonry -Advances in Non-linear Finite Element Methods. B.H.V. Topping and M. Papadrakakis(eds), Civil-Comp Press, Edinburgh, pp 143-157.

Pegon, P. and Pinto, A. V. (1996) - Seismic Study of Monumental Structures - Structural Anal-ysis, Modelling and Definition of Experimental Model - Report EUR No.16387 EN.ISIS, EC, JRC, Ispra (VA), Italy.

Page 440: Seismic Assessment of Reinforced Concrete Frame Structures

416 REFERENCES

Pinto, A. V. (1997) - Seismic Performance of Structures - Safety and Economical Aspects -Ph.D. Thesis, Instituto Superior Técnico, Lisboa, (in preparation).

Pinto, A. V. and Pegon, P. (1991) - Numerical Representation of Seismic Input Motion -Experimental and Numerical Methods in Earthquake Engineering. J. Donea & P.M.Jones eds. CEC, JRC, Ispra (VA), Italy.

Pinto, A. V. and Calvi, M. (1996) - Ongoing and Future Research in Support of Eurocode 8 -Proc. of the 11th World Conference on Earthquake Engineering, Acapulco, Mexico.

Pinto, A. V.; Negro, P.; Pegon, P. and Arêde, A. (1994) - Analysis of the 4-Storey R/C Build-ing to be Tested in the ELSA Reaction-wall Facility - Proc. of the 10th Conference onEarthquake Engineering, Vienna, Austria.

Pinto, A. V.; Verzeletti, G.; Pegon, P.; Magonette, G. E.; Negro, P. and Guedes, J. (1996) -Pseudo-Dynamic Testing of Large-Scale R/C Bridges in ELSA - Report EURNo.16378 EN, ISIS, EC, JRC, Ispra (VA), Italy.

Pipa, M. (1993) - Ductilidade de Elementos de Betão Armado Sujeitos a Acções Cíclicas -Influência das Características Mecânicas das Armaduras - Dissertação para Douto-ramento em Engenharia Civil, Instituto Superior Técnico, Lisboa, (in Portuguese).

Pires F. (1990) - Influência das Paredes de Alvenaria no Comportamento de Estruturas Reticu-ladas de Betão Armado Sujeitas a Acções Horizontais - Internal Report, LNEC, Lis-boa, (in Portuguese).

Porter, F. and Powell, G. H. (1971) - Static and Dynamic Analysis of Inelastic Frame Struc-tures - Report EERC 82-22, Earthquake Engineering Research Center, University ofCalifornia, Berkeley.

Póvoas, R. F. (1991) - Modelos Não-Lineares de Análise e Dimensionamento de EstruturasLaminares de Betão Armado Incluindo Efeitos Diferidos - Dissertação para Douto-ramento em Engenharia Civil, Faculdade de Engenharia da Universidade do Porto,Porto (in Portuguese).

R

Ramberg, R. and Osgood, W. R. (1943) - Description of Stress Strain Curve by Three Parame-ters - Technical Note 902, N.A.C.A.

REBAP - Regulamento de Estruturas de Betão Armado e Pré-esforçado (1984), ImprensaNacional Casa da Moeda, Lisboa, (in Portuguese).

Roufaiel, M.S.L. and Meyer, C. (1983) - Analysis of Damaged Concrete Frame Buildings -Report No. SF-PER-7924695-CU-I, Dep. of Civil Engineering, Columbia University,New York, N.Y.

Roufaiel, M. S. L. and Meyer, C. (1987) - Analytical Modelling of Hysteretic Behaviour of

Page 441: Seismic Assessment of Reinforced Concrete Frame Structures

REFERENCES 417

Reinforced Concrete Frames - Journal of Structural Engineering, ASCE, Vol. 113, No.3, pp. 429-444.

S

Shing, P. B.; Vannan, M. T. and Cater, E. (1990) - Implicit Time Integration for Pseudo-Dynamic Tests - Earthquake Engineering and Structural Dynamics, Vol.20, pp. 551-576.

Simons, J. W. and Powell, G. H. (1982) - Solution Strategies for Statically Non-linear Struc-tures - Report EERC 82-22, Earthquake Engineering Research Center, University ofCalifornia, Berkeley.

Soleimani, D.; Popov, E. P. and Bertero, V. V. (1979) - Nonlinear Beam Model for RC FrameAnalysis - 7th Conference on Electronic Computation, ASCE, St Louis, Missouri.

Spacone, E.; Ciampi, V. and Filippou, F. C. (1992) - A Beam Element for Seismic DamageAnalysis - Report EERC 92-07, Earthquake Engineering Research Center, Universityof California, Berkeley.

Spacone, E. (1994) - Flexibility Based Finite Element Models for the Non-linear Static andDynamic Analysis of Concrete Frame Structures - Ph.D. Thesis, Department of CivilEngineering, University of California, Berkeley.

Spacone, E.; Ciampi, V. and Filippou, F. C. (1996) - Mixed Formulation of Non-linear BeamFinite Element - Computers and Structures, Vol. 58, No. 1, pp. 71-83.

Spiegel, M. R. (1970) - Manual de Formulas y Tablas Matematicas - McGraw-Hill.

T

Takayanagi, T. and Schnobrich, W. C. (1976) - Computed Behaviour of Reinforced ConcreteCoupled Shear Walls - Structural Research Series No 434, Civil Engineering Studies,University of Illinois, Urbana.

Takayanagi, T. and Schnobrich, W. C. (1979) - Non-Linear Analysis of Coupled Wall Systems- Earthquake Engineering and Structural Dynamics, Vol. 7, pp. 1-22.

Takeda, T.; Sozen, M. A. and Nielsen, M. N. (1970) - Reinforced Concrete Response to Simu-lated Earthquakes - Journal of the Structural Division, ASCE, Vol. 96, ST12, pp. 2557-2573.

Taucer, F.; Spacone, E. and Filippou, F.C. (1991) - A Fiber Beam-Column Element for Seis-mic Response Analysis of R/C Structures - Report EERC 91-17, Earthquake Engineer-ing Research Center, University of California, Berkeley.

Page 442: Seismic Assessment of Reinforced Concrete Frame Structures

418 REFERENCES

Thewalt, C. R. and Mahin, S. A. (1991) - The Pseudo-Dynamic Test Method: Verification andExtensions - Experimental and Numerical Methods in Earthquake Engineering. J.Donea & P.M. Jones eds. CEC, JRC, Ispra (VA), Italy.

Tjebbes, M.R. (1994) - Determination of the Equivalent Slab Width from Experimental Data -Special Publication No.I.94.139. STI, CEC, JRC, Ispra (VA), Italy.

U

Uang, C.-M. and Bertero, V.V. (1988) - Use of Energy as a Design Criterion in Earthquake-Resistant Design - Report EERC 88-18, Earthquake Engineering Research Center, Uni-versity of California, Berkeley.

V

Vaz, P.T. (1992) - Comportamento Sísmico de Pontes com Pilares de Betão Armado - Verifi-cação de Segurança - Dissertação para Doutoramento em Engenharia Civil, Faculdadede Engenharia da Universidade do Porto, Porto (in Portuguese).

W

Wen, Y. K. (1980) - Equivalent Linearization for Hysteretic Systems Under Random Excita-tion - Journal of Applied Mechanics, ASME, Vol. 47, No. 1, pp. 150-154.

Z

Zeris, C. A. (1986) - Three Dimensional Nonlinear Response of Reinforced Concrete Build-ings - Ph.D. Thesis, Department of Civil Engineering, University of California, Berke-ley.

Zeris, C. A. and Mahin, S. A. (1988) - Analysis of Reinforced Concrete Beam-Columns underUniaxial Excitation - Journal of Structural Engineering, ASCE, Vol. 114, ST4, pp.804-820.

Zienkiewicz, O. C. and Taylor, R. L. (1989) - The Finite Element Method. Volume 1. BasicFormulation and Linear Problems - Fourth Edition, McGraw Hill, London.

Page 443: Seismic Assessment of Reinforced Concrete Frame Structures

APPENDICES

Page 444: Seismic Assessment of Reinforced Concrete Frame Structures

Appendix A

Linear Elastic Timoshenko Beam Formulation

A.1 Section formulation and constitutive relationship

The main expressions of the Timoshenko beam theory are briefly recalled in this appendix.

Detailing of this well known formulation is deemed unnecessary in the present work, since it

can be easily found in the literature.

Notation and conventions are the same as introduced in 2.4 and the bases for theory derivation

and application are adopted as used in Pegon (1993). Hence, referring to Figure 2.5, it is

assumed that:

• the element reference system (x,y,z) consists of the principal inertia axes and the same

applies to the local section reference frame;

• translation components of the displacement vector refer to the section centroid and

rotation components refer to the section plane.

Thus, according to the Timoshenko formulation accounting for shear deformations, the com-

ponents of the generalized strain vector are given by

(A.1)

which, for the elastic case, are related with the generalized stress vector components by

(A.2)

a x( )

e x( )

εx xddux=

ϕx xddθx=

βy xdduy θz–=

ϕy xddθy=

βz xdduz θy+=

ϕz xddθz=

S x( )

εx Nx EA( )⁄=

ϕx Mx GJx( )⁄=

βy Vy GAy( )⁄=

ϕy My EIy( )⁄=

βz Vz GAz( )⁄=

ϕz Mz EIz( )⁄=

Page 445: Seismic Assessment of Reinforced Concrete Frame Structures

422 Appendix A

where:

• A, and stand, respectively, for the cross-sectional area and the shear reduced areas in

both y and z directions;

• , and refer to the modified moment of polar inertia and to the moments of flexural

inertia in the y and z directions, respectively;

• E and G are the Young and distortion modulus.

All these geometric and mechanical characteristics, may vary along the element but, as for sec-

tion deformations and forces, the reference to abscissa x was suppressed for simplicity.

, and are introduced in Eq. (A.2) to account for possible cross-sectional warping, and

are usually given by

(A.3)

where and are parameters less then or equal to 1, depending on the Poisson ratio and on

the cross-section shape and dimensions (Dias da Silva (1995)).

Therefore, the section flexibility matrix can be written in the diagonal form

(A.4)

whose terms can be uniform or vary along the element according to an assumed distribution.

A.2 Element flexibility matrix

In the present work, the element is considered divided into two parts, along which the cross-

sectional properties are assumed uniform, as shown in Figure A.1; the subscript or

is assigned to properties in correspondence with the respective element part. Thus, the

flexibility distribution is defined by

(A.5)

where h is the relative abscissa of the span section H and both and are diagonal

Ay Az

Jx Iy Iz

Ay Az Jx

Ay αyA= Az αzA= Jx αyIy αzIz+=

αy αz

f x( ) diag EA( )-1 GAy( )-1 GAz( )-1 GJx( )-1 EIy( )-1 EIz( )-1, , , , ,[ ]=

i 1=

i 2=

f x( )f1 x( )

f2 x( )⎩⎨⎧

=forfor

0 x hL≤ ≤hL x L≤ ≤

f1 x( ) f2 x( )

Page 446: Seismic Assessment of Reinforced Concrete Frame Structures

Linear Elastic Timoshenko Beam Formulation 423

matrices of constant terms given by

(A.6)

Figure A.1 Distribution of flexibility properties along the element

The element flexibility matrix is obtained by integration of according to Eq. (2.25);

however, a slight modification is introduced in the notation of the matrix, in order to

highlight partial contributions to .

Without considering element applied loads (irrelevant for the present purposes), Eq. (2.14) can

be re-written as

(A.7)

where the row-matrices stand for each internal force component (indicated as super-

script) and are given by

(A.8)

Due to the diagonal nature of , the integral of Eq. (2.25) to obtain , leads to the follow-

ing matrix summation

fi x( ) diag EA( )i-1 GAy( )i

-1 GAz( )i-1 GJx( )i

-1 EIy( )i-1 EIz( )i

-1, , , , ,[ ]=

E1 E2HxH hL=

x

f2 x( )f1 x( )

F f x( )

b x( )

F

Nx

Vy

Vz

Mx

My

Mz⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

x( )

bNx x( )

bVy x( )

bVz x( )

bMx x( )

bMy x( )

bMz x( )

Fx1

Mx1

My1

Mz1

My2

Mz2⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

⋅=

b… x( )

bNx x( ) 1– 0 0 0 0 0=

bVy x( ) 0 0 0 1 L⁄– 0 1 L⁄–=

bVz x( ) 0 0 1 L⁄ 0 1 L⁄ 0=

bMx x( ) 0 1– 0 0 0 0=

bMy x( ) 0 0 x L⁄ 1–( ) 0 x L⁄ 0=

bMz x( ) 0 0 0 x L⁄ 1–( ) 0 x L⁄=

f x( ) F

Page 447: Seismic Assessment of Reinforced Concrete Frame Structures

424 Appendix A

(A.9)

each one accounting for the contribution of the internal force/deformation component indi-

cated by the superscript. The respective matrix expressions are given by

(A.10)

where the same superscript has also been included in the section flexibility terms.

Using the flexibility distribution given by Eq. (A.5) and the expressions (A.8), the six flexibil-

ity matrix contributions are written as:

(A.11)

(A.12)

(A.13)

(A.14)

F FNx F

Vy FVz F

Mx FMy F

Mz+ + + + +=

FNx b

Nx x( )[ ]T

fNx x( ) b

Nx x( )⋅ ⋅ xd0

L

∫=

FVy b

Vy x( )[ ]T

fVy x( ) b

Vy x( )⋅ ⋅ xd0

L

∫=

FVz b

Vz x( )[ ]T

fVz x( ) b

Vz x( )⋅ ⋅ xd0

L

∫=

FMx b

Mx x( )[ ]T

fMx x( ) b

Mx x( )⋅ ⋅ xd0

L

∫=

FMy b

My x( )[ ]T

fMy x( ) b

My x( )⋅ ⋅ xd0

L

∫=

FMz b

Mz x( )[ ]T

fMz x( ) b

Mz x( )⋅ ⋅ xd0

L

∫=

FNx

F11Nx L

EA( )1---------------h L

EA( )2--------------- 1 h–( )+=

FlmNx 0= for l 1≠( ) or m 1≠( )⎩

⎪⎨⎪⎧

= FMx

F22Mx L

GJx( )1-----------------h L

GJx( )2----------------- 1 h–( )+=

FlmMx 0= for l 2≠( ) or m 2≠( )⎩

⎪⎨⎪⎧

=

FVy

FlmVy L 1–

GAy( )1------------------h L 1–

GAy( )2------------------ 1 h–( )+= for l 4 6,=( ) and m 4 6,=( )

FlmVy 0= for l 4 6,≠( ) or m 4 6,≠( )⎩

⎪⎨⎪⎧

=

FVz

FlmVz L 1–

GAz( )1-----------------h L 1–

GAz( )2----------------- 1 h–( )+= for l 3 5,=( ) and m 3 5,=( )

FlmVz 0= for l 3 5,≠( ) or m 3 5,≠( )⎩

⎪⎨⎪⎧

=

FMy

F33My L

EIy( )1---------------ξa1

LEIy( )2

---------------ξa2+=

F35My F53

My LEIy( )1

---------------ξb1

LEIy( )2

---------------ξb2+= =

F55My L

EIy( )1---------------ξc1

LEIy( )2

---------------ξc2+=

FlmMy 0= for l 3 5,≠( ) or m 3 5,≠( )⎩

⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎧

=

Page 448: Seismic Assessment of Reinforced Concrete Frame Structures

Linear Elastic Timoshenko Beam Formulation 425

(A.15)

where the coefficients , , and (for ), depend only on the element subdivision

factor h and are given by

(A.16)

For the particular case of uniform properties along the whole element ( )

any value of h is suitable because only the sum of each pair of coefficients affecting section

flexibility terms is relevant and it remains constant regardless of h.

Therefore, after summation of all contributions, the total flexibility matrix becomes

(A.17)

highlighting that force/displacement component interaction exists only between those contrib-

uting for the same direction of deformation (namely and , or and ).

FMz

F44Mz L

EIz( )1---------------ξa1

LEIz( )2

---------------ξa2+=

F46Mz F64

Mz LEIz( )1

---------------ξb1

LEIz( )2

---------------ξb2+= =

F66Mz L

EIz( )1---------------ξc1

LEIz( )2

---------------ξc2+=

FlmMz 0= for l 4 6,≠( ) or m 4 6,≠( )⎩

⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎧

=

ξaiξbi

ξcii 1 2,=

ξa1

h3

3----- h2 h–( )–=

ξb1

h3

3----- h2

2-----–=

ξc1

h3

3-----=

ξa2

1 h3–3

-------------- h2 h–( )+ 13--- ξa1

–= =

ξb2

1 h3–3

-------------- 1 h2–2

--------------– 16---– ξb1

–= =

ξc2

1 h3–3

-------------- 13--- ξc1

–= =

f x( ) f1 x( ) f2 x( )= =

F

F

LEA------- 0 0 0 0 0

0 LGJx--------- 0 0 0 0

0 0 L 1–

GAz---------- L

3EIy-----------+⎝ ⎠

⎛ ⎞ 0 L 1–

GAz---------- L

6EIy-----------–⎝ ⎠

⎛ ⎞ 0

0 0 0 L 1–

GAy---------- L

3EIz-----------+⎝ ⎠

⎛ ⎞ 0 L 1–

GAy---------- L

6EIz-----------–⎝ ⎠

⎛ ⎞

0 0 L 1–

GAz---------- L

6EIy-----------–⎝ ⎠

⎛ ⎞ 0 L 1–

GAz---------- L

3EIy-----------+⎝ ⎠

⎛ ⎞ 0

0 0 0 L 1–

GAy---------- L

6EIz-----------–⎝ ⎠

⎛ ⎞ 0 L 1–

GAy---------- L

3EIz-----------+⎝ ⎠

⎛ ⎞

=

My1My2

Mz1Mz2

Page 449: Seismic Assessment of Reinforced Concrete Frame Structures

426 Appendix A

A.3 Element displacements as integrated deformations

The element displacement vector , given by Eq. (2.16) as an integral of deformations, is

detailed next, for the case of element applied forces in three directions. Expressions are

obtained for the element axis system (x,y,z); therefore, the vector given by Eq. (C.3) in

Appendix C must be adopted in the following.

According to Eq. (2.14), and making use of , the elastic section deformations are given by

(A.18)

which, after substitution in Eq. (2.16), lead to , where stands for displace-

ments due to element applied loads, expressed by

(A.19)

Actually, only needs to be detailed, since the flexibility matrix is already known. Fol-

lowing a similar procedure to that used in A.2, can also be decomposed into the contribu-

tions arising from the distinct force/deformation components.

Thus, using the row-matrices, expressed by Eqs. (A.8), and substituting

by its components, Eq. (A.19) yields

(A.20)

where, again, the superscript identifies the contributing components. These are given by

(A.21)

u

Sp x( )

f x( )

e x( ) f x( ) b x( ) Q⋅ ⋅ f x( ) Sp x( )⋅+=

u F Q⋅ up+= up

up bT x( ) f x( ) Sp x( )⋅ ⋅ xd0

L

∫=

up F

up

b… x( )

ep x( ) f x( ) Sp x( )⋅=

up upNx up

Vy upVz up

Mx upMy up

Mz+ + + + +=

upNx

εx( )p– xd0

L

∫00000⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= upMx

0

ϕx( )p– xd0

L

∫0000⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

Page 450: Seismic Assessment of Reinforced Concrete Frame Structures

Linear Elastic Timoshenko Beam Formulation 427

(A.22)

(A.23)

whose deformation components inside the integrals are related to internal forces by Eq. (A.2).

For the flexibility distribution given by Eq. (A.5) and the vector components (see Eq.

(C.3)), the previous integrals yield the following expressions

(A.24)

(A.25)

upVz

00βz( )pL

------------ xd0

L

∫0βz( )pL

------------ xd0

L

∫0⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= upMy

00

xL--- 1–⎝ ⎠⎛ ⎞ ϕy( )p xd

0

L

∫0

xL---⎝ ⎠⎛ ⎞ ϕy( )p xd

0

L

∫0⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

upVy

000βy( )p–L

---------------- xd0

L

∫0βy( )p–L

---------------- xd0

L

∫⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= upMz

000

xL--- 1–⎝ ⎠⎛ ⎞ ϕz( )p xd

0

L

∫0

xL---⎝ ⎠⎛ ⎞ ϕz( )p xd

0

L

∫⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

Sp x( )

εx( )p– xd0

L

∫pxL2

2----------- h2

EA( )1--------------- 1 h2–( )

EA( )2-------------------+ PxL 1 h–

EA( )2---------------+=

ϕx( )p– xd0

L

∫ 0=

βy( )p–L

---------------- xd0

L

∫p– yL2

------------ηs1

GAy( )1------------------

ηs2

GAy( )2------------------+ Py

ηs1

GAy( )1------------------

ηs2

GAy( )2------------------+–=

βz( )pL

------------ xd0

L

∫pzL2

--------ηs1

GAz( )1-----------------

ηs2

GAz( )2-----------------+ Pz

ηs1

GAz( )1-----------------

ηs2

GAz( )2-----------------++=

xL--- 1–⎝ ⎠⎛ ⎞ ϕy( )p xd

0

L

∫ p– zL3 ηa1

EIy( )1---------------

ηa2

EIy( )2---------------+ PzL

2 ηc1

EIy( )1---------------

ηc2

EIy( )2---------------+–=

xL---⎝ ⎠⎛ ⎞ ϕy( )p xd

0

L

∫ p– zL3 ηb1

EIy( )1---------------

ηb2

EIy( )2---------------+ PzL2 ηd1

EIy( )1---------------

ηd2

EIy( )2---------------+–=

xL--- 1–⎝ ⎠⎛ ⎞ ϕz( )p xd

0

L

∫ pyL3 ηa1

EIz( )1---------------

ηa2

EIz( )2---------------+ PyL2 ηc1

EIz( )1---------------

ηc2

EIz( )2---------------++=

xL---⎝ ⎠⎛ ⎞ ϕz( )p xd

0

L

∫ pyL3 ηb1

EIz( )1---------------

ηb2

EIz( )2---------------+ PyL2 ηd1

EIz( )1---------------

ηd2

EIz( )2---------------++=

Page 451: Seismic Assessment of Reinforced Concrete Frame Structures

428 Appendix A

where the coefficients , , , and for are given by

(A.26)

with , and as expressed by Eq. (A.16). For the case of element uniform properties

with a possible concentrated force at H (thus, ), these coefficients reduce to

Adopting the following weighted average flexibilities

(A.27)

where the subscripts p (lower case) and P (upper case) stand, respectively, for the distributed

and the concentrated force contributions, the vector becomes

i

1 7/24 -1/12 1/24 1/4 11/384 -5/384 1/24 -1/48

2 1/24 -1/12 7/24 -1/4 5/384 -11/384 1/48 -1/24

1/3 -1/6 1/3 0 1/24 -1/24 1/16 -1/16

ηsiηai

ηbiηci

ηdii 1 2,=

ηs1h h2–=

ηa1h4 8⁄ h3 3⁄ h2 4⁄+–=

ηb1h4 8⁄ h3 6⁄–=

ηc11 h–( ) ξb1

–( )=

ηd11 h–( ) ξc1

–( )=

ηs2ηs1

–=

ηa21 24⁄ ηa1

–=

ηb21 24⁄– ηb1

–=

ηc2h 1 3⁄ ξa1

–( ) hξa2= =

ηd2h 1 6⁄– ξb1

–( ) hξb2= =

ξaiξbi

ξci

h 1 2⁄=

ξaiξbi

ξciηsi

ηaiηbi

ηciηdi

Σi

1EA( )p

--------------- h2

EA( )1--------------- 1 h2–( )

EA( )2-------------------+=

1GAy

----------ηs1

GAy( )1------------------

ηs2

GAy( )2------------------+=

1EIy( )p1

-----------------ηa1

EIy( )1---------------

ηa2

EIy( )2---------------+=

1EIy( )p2

-----------------ηb1

EIy( )1---------------

ηb2

EIy( )2---------------+=

1EIz( )p1

-----------------ηa1

EIz( )1---------------

ηa2

EIz( )2---------------+=

1EIz( )p2

-----------------ηb1

EIz( )1---------------

ηb2

EIz( )2---------------+=

1EA( )P

--------------- 1 h–EA( )2

---------------=

1GAz

----------ηs1

GAz( )1-----------------

ηs2

GAz( )2-----------------+=

1EIy( )P1

------------------ηc1

EIy( )1---------------

ηc2

EIy( )2---------------+=

1EIy( )P2

------------------ηd1

EIy( )1---------------

ηd2

EIy( )2---------------+=

1EIz( )P1

-----------------ηc1

EIz( )1---------------

ηc2

EIz( )2---------------+=

1EIz( )P2

-----------------ηd1

EIz( )1---------------

ηd2

EIz( )2---------------+=

up

Page 452: Seismic Assessment of Reinforced Concrete Frame Structures

Linear Elastic Timoshenko Beam Formulation 429

(A.28)

For the particular case of uniform element properties and this simplifies to

(A.29)

showing that the shear contributions to rotations vanish due to anti-symmetry of shear distor-

tion diagrams along the element.

Additionally, this particular case helps to estimate the concentrated load at H equivalent to the

distributed one as referred in 3.4. If the equivalence criterion consists in the equality of elastic

fixed end section moments, it is identically enforced if the equality of elastic end section dis-

placements is imposed. In this context, if a “fictitious” concentrated load vector Peq is sought,

equivalent to the “real” distributed one p, the components of Peq must be such that the right-

most terms of components in Eq. (A.29) be equal to the leftmost ones. Therefore, the fol-

lowing equivalence conditions are obtained

up

ux1( )pθx1( )pθy1( )pθz1( )pθy2( )pθz2( )p⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

pxL2

2 EA( )p------------------ Px

LEA( )P

---------------+

0

pzL

2GAz

------------- L3

EIy( )p1

-----------------– Pz1

GAz

---------- L2

EIy( )P1

------------------–+

p– yL

2GAy

-------------- L3

EIz( )p1

-----------------– Py1

GAy

---------- L2

EIz( )P1

-----------------––

pzL

2GAz

------------- L3

EIy( )p2

-----------------– Pz1

GAz

---------- L2

EIy( )P2

------------------–+

p– yL

2GAy

-------------- L3

EIz( )p2

-----------------– Py1

GAy

---------- L2

EIz( )P2

-----------------––⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= =

h 1 2⁄=

up

pxL2

2EA-----------⎝ ⎠⎛ ⎞ Px

L2EA-----------⎝ ⎠⎛ ⎞+

0

pzL– 3

24EIy--------------⎝ ⎠⎜ ⎟⎛ ⎞

PzL– 2

16EIy--------------⎝ ⎠⎜ ⎟⎛ ⎞

+

pyL3

24EIz--------------⎝ ⎠⎜ ⎟⎛ ⎞

PyL2

16EIz--------------⎝ ⎠⎜ ⎟⎛ ⎞

+

pzL3

24EIy--------------⎝ ⎠⎜ ⎟⎛ ⎞

PzL2

16EIy--------------⎝ ⎠⎜ ⎟⎛ ⎞

+

pyL– 3

24EIz--------------⎝ ⎠⎜ ⎟⎛ ⎞

PyL– 2

16EIz--------------⎝ ⎠⎜ ⎟⎛ ⎞

+⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

up

Page 453: Seismic Assessment of Reinforced Concrete Frame Structures

430 Appendix A

(A.30)

which have been adopted in the present study for the approximate consideration of distributed

forces.

Px( )eq pxL=

Py( )eq23---p

yL=

Pz( )eq23---p

zL=

Page 454: Seismic Assessment of Reinforced Concrete Frame Structures

Appendix B

Trilinear Model Details

In this appendix some details are included about of the adopted trilinear model. The primary

curve (see Figure 3.2) and some basic definitions were already introduced in 3.3; the hysteretic

behaviour rules are described in the next paragraphs.

A general loading path is illustrated in Figure B.1 to exemplify typical features of the hystere-

tic process; neither pinching nor strength degradation effects are included in this case.

Figure B.1 Hysteretic rules of the trilinear model. General loading path

5

1

3

4

2

8

6

6

9

77

8

a

e , k

d

c

b , g

f

l

j

i , oh

n

m

+

+

+

+

+

+

+

+

++

++

My+

Mc-

My-

Mc+

α-My-

α+My+

ϕy-

ϕy+

ϕ

M

A+

A-

9

E+ ME+ ϕE

+( , )

E- ME- ϕE

-( , )

Loading path:

a-b-c-d-e-f-g-h-i-j-k-l-m-n-o

k1+

k1-

k2+

k2-

1st 2nd 3rd Cycle:

Page 455: Seismic Assessment of Reinforced Concrete Frame Structures

432 Appendix B

Branches 1 to 5 refer to loading situations; the unloading lines are labelled with 6 and 7, while

the reloading lines are identified by 8 and 9.

Unloading lines are obtained so as to induce stiffness degradation for increasing inelastic

deformation; for this purpose, a common target point is considered ( or ) in the extrapo-

lated elastic line at a moment level or , where and control the degree of

unloading stiffness degradation. Unloading lines aim at this target point and keep this stiffness

until they reach the axis. It is apparent that low values of increase the stiffness degrada-

tion and, consequently, reduce the area enclosed by the hysteresis loops and the corresponding

dissipated energy.

Reloading branches 8 and 9, included in Figure B.1, aim at target points and , respec-

tively, where the maximum absolute value of deformation has been found in previous cycles

(for the same sign of moments).

Although these are the general rules for unloading and reloading, some special situations are

considered:

• if the section has cracked for one direction of bending, but still remains uncracked for the

other direction, then, when reloading, the target is the cracking point on that side, as is the

case of path a-b-c-d; then, loading follows along the cracked branch (2 or 3);

• in the case of unloading or reloading for interior cycles, i.e. enclosed between the outermost

lines as is the case in Figure B.2, the stiffness remains “frozen” at that of the corresponding

outermost lines (for example, the unloading branches 12 and 14 are parallel to line 6, and

the same happens between lines 13, 15 and 7); although not labelled, the positive and nega-

tive interior reloading branches are parallel to the lines 9 and 8, respectively.

The pinching effect is given by a reduction of the hysteresis loop area caused by a low reload-

ing stiffness that may increase after a certain amount of plastic deformation is recovered. The

effect is mainly due to development of cracks that remain open during almost the entire cycle;

in such case, some phases of the loading sequence are very likely to exhibit a rather low stiff-

ness, since the applied moment is only resisted by tensile and compressive forces in the steel.

Stiffness recovering may occur if further deformations go far enough to close the opened

cracks and to engage again the stiffness of the compressed concrete area.

A+ A-

α+My+

– α-My-

– α+ α-

ϕ α

E- E+

Page 456: Seismic Assessment of Reinforced Concrete Frame Structures

Trilinear Model Details 433

Figure B.2 Trilinear model. The pinching effect and interior cycles

This phenomenon is commonly associated with pronounced asymmetry of top and bottom

reinforcement, but other aspects may contribute to the increase of the pinching effect, such as:

• loss or accentuated deterioration of bond, inducing significant slip between concrete and

steel;

• shear deformations, with a high relative weight when compared to the flexural deforma-

tions, causing important shear stress to be transferred by aggregate interlock and dowel

action (thus, with a very low stiffness) while cracks remain open.

Modelling of pinching, when it is to be included, is performed by lowering the target point of

reloading branches immediately after crossing the axis. Instead of the point E (see Figure

B.2), lines 8 (or 9) aim at point B existing on the previous unloading line at a level (or

), until reaching the so called crack-closing deformation (or ) assumed correspond-

ing to null moment in the previous unloading phase. After this point (or ) new reloading

lines 10 or 11 aim again at the previous maximum point (or ) and the process goes on as

for the case of no pinching.

Strength degradation is likely to develop in RC sections subjected to cyclic loading, mainly

due to increasing deterioration of the concrete resisting capacity. The closing process of open

Ms- γ-My

-=

ϕ

M

ϕs-

1k1

+

ϕs+

C+

Y+

C-

Y-

B+

E+

E-

k3+

k2+

k2-

k1-

k3-

B-

9

8

6

713

11

10

12

14

15

ϕx-

ϕx+

X+

X-

Ms+ γ+My

+=

ϕ

γ+My+

γ-My- ϕx

+ ϕx-

X+ X-

E+ E-

Page 457: Seismic Assessment of Reinforced Concrete Frame Structures

434 Appendix B

cracks becomes less effective during cyclic loading, since the roughness of the contact surface

progressively increases. Moreover, steel-concrete bond also deteriorates along reinforcing

bars, at increasing distances from the crack lips, thus reducing the effectiveness of stress trans-

fer between steel and the surrounding concrete due to lack of local anchorage. These two

aspects lead to both stiffness and strength reduction for increasing deformations and number of

cycles, and is often referred to as softening effect.

In the present model this effect is taken into account by means of the dissipated hysteretic

energy contribution to the increase of section deformations (Kunnath et al. (1990)); the sche-

matic process is shown in Figure B.3.

Figure B.3 Trilinear model. Strength deterioration rule

The parameter is defined as the ratio of incremental damage caused by the increase of max-

imum deformation to the normalized increment of hysteretic energy, , responsible for

such deformation increase. It is given, in terms of moment and curvatures, by

(B.1)

where is the ultimate curvature under monotonic loading. The energy increment corre-

sponds to the hysteretic dissipation occurred in the cycle starting at the previous maximum

ϕ

M

C+

Y+

C-

Y-

E+

E-

∆ϕ+ β+ ∆E+

My+

----------⎝ ⎠⎜ ⎟⎛ ⎞

=

∆E+ (Dissipated hysteretic energy

T+ MT+ ϕT

+( , )

in the cycle starting at E+)

Prediction

Predictionpoint

reloading line

β

dϕm dE

βdϕm

ϕu----------⎝ ⎠

⎛ ⎞ dEϕuMy-------------⎝ ⎠

⎛ ⎞⁄dϕmMy

dE-----------------= =

ϕu dE

Page 458: Seismic Assessment of Reinforced Concrete Frame Structures

Trilinear Model Details 435

deformation , which is likely to include smaller internal loops also contributing to . In

fact this parameter appears in the well known Park and Ang damage index definition (Park

et al. (1984)), weighting the dissipated energy contribution to the total damage.

Using Eq. (B.1) the deformation increment can be obtained by

(B.2)

and, obviously, it is to be computed for both directions of loading.

If no pinching effect is considered, the reloading line at the prediction point aims at the previ-

ous maximum point ; this allows for an estimate of the hysteretic dissipated energy in

that cycle (including internal ones). Thereby, the increment of curvature can be computed

by Eq. (B.2) and a modified target point is defined on the primary curve at curvature

. The actual reloading branch stiffness is then obtained aiming at this new target

point. If pinching is also to be included, the procedure is affected in the same way, but only for

the second reloading branch, i.e. for the stiffness determination after the point (see Fig-

ure B.2).

It must be emphasized that this strength degradation procedure does not exactly correspond to

the softening effect, since it does not ensure an effective reduction of strength for increasing

deformation. In fact, with this methodology a strength drop is achieved between successive

cycles at the same maximum deformation, but for increasing deformations the strength is not

prevented to increase because there is no explicit softening branch in the primary curve. There-

fore, if the modelling of softening is thought to be important, such as in the case of sections

highly controlled by the compressed concrete behaviour (e.g. heavily loaded columns), another

descending branch should be included in the primary curve after the peak moment is reached.

Concerning the values of hysteretic model parameters, the following suggestions can be found

in Kunnath et al. (1990):

• For the unloading stiffness degradation, values of between 2 and 4 appear to give good

results for well detailed RC sections and the peak response is not very sensitive to values.

Note that with an origin oriented model is obtained, while setting a non-

degrading Clough-type model is achieved.

ϕm dE

β

dϕm βdEMy-------=

E+ ∆E+

∆ϕ

T+

ϕE+ ∆ϕ++

k3+ X+

α

α

α 0= α ∞=

Page 459: Seismic Assessment of Reinforced Concrete Frame Structures

436 Appendix B

• The pinching behaviour due to top and bottom reinforcement asymmetry (such as the beam

cases) is well simulated when the pinching moment in the stronger side has approxi-

mately the same magnitude of the weaker side yielding moment. This means that, when no

other causes for pinching are present, values can be given by

(B.3)

which implicitly assumes that pinching only occurs below the yielding moment. If other

pinching sources are to be included (such as bond-slip or high shear deformations) then

other values have to be selected depending upon the degree of pinching expected. The

authors refer that low values in the range of 0.01 to 0.1 are not unusual for cases with high

contribution of shear deformation.

• For the strength deterioration parameter an empirical expression, based on a great

amount of experimental results, is commonly used (Kunnath et al. (1990)) and given (with

slight algebraic adjustments) by

(B.4)

where ρw is the volumetric confinement ratio (i.e. the volume of closed stirrups divided by

the volume of confined concrete core), ν is the normalized axial force (taken positive if

compressive) and ωt is the mechanical ratio of tension reinforcement; typical values of

are not far from 0.05.

This model description did not intend to be exhaustive. However it is believed that the above

paragraphs and figures help to understand the main features, capabilities and limitations of the

model, which is important for the interpretation of the numerical results presented in this work.

Ms

γ

γ+ My-

My+

------- 1≤= γ- My+

My-

------- 1≤=

γ

β( )

β 0.9100ρw 0.37max ν 0.05( , ) 0.5 ωt 0.17–( )2+( )=

β

Page 460: Seismic Assessment of Reinforced Concrete Frame Structures

Appendix C

Internal Force Distributions

C.1 General

In this appendix, several expressions are included for internal force distribution due to both

uniformly distributed forces applied along the element and concentrated ones applied in the

span section H; only forces are considered, i.e. applied moments are not included.

The general case of non-zero force components in three directions is first introduced in C.2 and

the particular case where element loads exist only in the non-linear bending plane is further

detailed in C.3. The corresponding expressions for abscissas of internal moving sections and

for their derivatives with respect to applied moments are included in C.4.

C.2 Element applied forces in three directions

Consider again Figure 2.6, now with uniformly distributed forces p in the whole flexible ele-

ment and concentrated forces P acting in section H. With the notation adopted there, and

denoting by the relative abscissa of H, the vector of resultants expressed by

Eq. (2.7) becomes

(C.1)

h xH L⁄= R x( )

R x hL<( )

x x( )

y x( )

z x( )

x x( )

y x( )

z x( )⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫ pxx

pyx

pzx0

pzx2

2-----

pyx2

2-----–

⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= = ; R x hL≥( )

pxx Px+

pyx Py+

pzx Pz+

0

pzx2

2----- Pz x hL–( )+

pyx2

2-----– Py x hL–( )–

⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

Page 461: Seismic Assessment of Reinforced Concrete Frame Structures

438 Appendix C

while the vector defined by Eq. (2.8) is obtained by substituting L for x in Eq. (C.1).

Thereby, the vector given by Eq. (2.12), referring to the element load contribution to the

forces, can be expressed by

(C.2)

and the contributions to section internal forces, expressed both in the element axes (x,y,z)

(given by , Eq. (2.15)) and in the local axis system (given by , Eq. (3.1)), are now

written, respectively, as

(C.3)

and

(C.4)

Therefore, the expression (3.5) of the generic section internal forces , in terms of the end

section ones , can be detailed as follows

RL

Qpf

Qf

Qpf

Qp1

f

Qp2

f⎩ ⎭⎪ ⎪⎨ ⎬⎪ ⎪⎧ ⎫

= with Qp1

f

0

p– yL2--- Py 1 h–( )–

p– zL2--- Pz 1 h–( )–

000⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= and Qp2

f

p– xL Px–

p– yL2--- Pyh–

p– zL2--- Pzh–

000⎩ ⎭

⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

Sp x( ) Spsx( )

Sp x hL<( )

p– xx

pyL2--- x–⎝ ⎠⎛ ⎞ Py 1 h–( )+

pzL2--- x–⎝ ⎠⎛ ⎞ Pz 1 h–( )+

0

pz L x–( )x2--- Pz 1 h–( )x+

p– y L x–( )x2--- Py 1 h–( )x–

⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= ; Sp x hL≥( )

p– xx Px–

pyL2--- x–⎝ ⎠⎛ ⎞ Pyh–

pzL2--- x–⎝ ⎠⎛ ⎞ Pzh–

0

pz L x–( )x2--- Pz L x–( )h+

p– y L x–( )x2--- Py L x–( )h–

⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

Spsx hL<( )

p– xx

p– yL2--- x–⎝ ⎠⎛ ⎞ Py 1 h–( )–

p– zL2--- x–⎝ ⎠⎛ ⎞ Pz 1 h–( )–

0

p– z L x–( )x2--- Pz 1 h–( )x–

py L x–( )x2--- Py 1 h–( )x+

⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

= ; Spsx hL≥( )

p– xx Px–

p– yL2--- x–⎝ ⎠⎛ ⎞ Pyh+

p– zL2--- x–⎝ ⎠⎛ ⎞ Pzh+

0

p– z L x–( )x2--- Pz L x–( )h–

py L x–( )x2--- Py L x–( )h+

⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

=

Ss x( )

Qes

Page 462: Seismic Assessment of Reinforced Concrete Frame Structures

Internal Force Distributions 439

(C.5)

and

(C.6)

where the superscripts and have been included in the components, in order to

remind that they consist of internal forces associated with the end section local axis systems;

Figure 3.1 helps to clarify the meaning of Eqs. (C.5) and (C.6).

C.3 Element applied forces only in the non-linear bending plane

For the particular case of element forces applied only in the non-linear bending plane, by

default xz, internal forces are readily obtained from Eqs. (C.5) and (C.6). The influence of ele-

ment loads appears only in the components and , which are likely to affect deforma-

tions in the xz plane. Due to their importance in the present work context, these components are

further detailed next, with a simplified notation: subscripts and are suppressed, the end

section moments of interest ( and ) are simply denoted by and , and the dis-

tributed and concentrated forces ( and ) are referred to as p and P, respectively.

NxsNx

E1 pxx–=

Vys

MzE1

L---------

MzE2

L---------– py

L2--- x–⎝ ⎠⎛ ⎞– Py 1 h–( )–=

Vzs

MyE1

L---------–

MyE2

L--------- pz

L2--- x–⎝ ⎠⎛ ⎞– Pz 1 h–( )–+=

MxsMx

E1=

Mys1 x

L---–⎝ ⎠

⎛ ⎞MyE1 x

L---⎝ ⎠⎛ ⎞My

E2 pz L x–( )x2---– Pz 1 h–( )x–+=

Mzs1 x

L---–⎝ ⎠

⎛ ⎞MzE1 x

L---⎝ ⎠⎛ ⎞Mz

E2 py L x–( )x2--- Py 1 h–( )x+ + +=⎝

⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎛

for x hL<( )

NxsNx

E1 pxx– Px–=

Vys

MzE1

L---------

MzE2

L---------– py

L2--- x–⎝ ⎠⎛ ⎞– Pyh+=

Vzs

MyE1

L---------–

MyE2

L--------- pz

L2--- x–⎝ ⎠⎛ ⎞– Pzh+ +=

MxsMx

E1=

Mys1 x

L---–⎝ ⎠

⎛ ⎞MyE1 x

L---⎝ ⎠⎛ ⎞My

E2 pz L x–( )x2---– Pz L x–( )h–+=

Mzs1 x

L---–⎝ ⎠

⎛ ⎞MzE1 x

L---⎝ ⎠⎛ ⎞Mz

E2 py L x–( )x2--- Py L x–( )h+ + +=⎝

⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎛

for x hL≥( )

E1 E2 Qes

VzsMys

ys zs

MyE1 My

E2 ME1ME2

pz Pz

Page 463: Seismic Assessment of Reinforced Concrete Frame Structures

440 Appendix C

Therefore, according to Figure C.1, only the following expressions are considered:

(C.7)

and

(C.8)

Figure C.1 Element applied loads in the non-linear bending plan; simplified notation

Considering the local abscissas and , the relative local abscissas of section

H are given, respectively, by

(C.9)

Introducing the transverse force , the distributions of V and M can be

also written as

(C.10)

and

V x( )ME1

L---------–

ME2

L--------- p L

2--- x–⎝ ⎠⎛ ⎞– P 1 h–( )–+=

M x( ) 1 xL---–⎝ ⎠

⎛ ⎞ME1

xL---⎝ ⎠⎛ ⎞ME2

p L x–( )x2---– P 1 h–( )x–+=

for x hL<( )

V x( )ME1

L---------–

ME2

L--------- p L

2--- x–⎝ ⎠⎛ ⎞– Ph+ +=

M x( ) 1 xL---–⎝ ⎠

⎛ ⎞ME1

xL---⎝ ⎠⎛ ⎞ME2

p L x–( )x2---– P L x–( )h–+=

for x hL≥( )

p

ME1 ME2

MV

xs1 s2

L

H

sH1sH2

P

s1 x= s2 L x–=

h1sH1

L------- h= = and h2

sH2

L------- 1 h1–= =

VE ME2ME1

–( ) L⁄=

V s1( ) VE p L2--- s1–⎝ ⎠⎛ ⎞ P 1 h1–( )+–=

M s1( ) ME1VEs1 p L s1–( )

s12---- P 1 h1–( )s1+–+=

Page 464: Seismic Assessment of Reinforced Concrete Frame Structures

Internal Force Distributions 441

(C.11)

for the element parts and , respectively. Eqs. (C.10) and (C.11) are very convenient

for locating internal moving sections and can be condensed in only one, as follows

(C.12)

where the sign coefficient is defined by

(C.13)

C.4 Moving section abscissas and respective derivatives

Abscissas

The location of sections associated to a specific moment are obtained according to Eq.

(3.11), in terms of the element abscissa x, or according to Eq. (3.12) referring to local abscissas

. Since is generally associated to each end section, the use of Eq. (C.12), as a particular

case of Eq. (3.12), is more adequate and, is therefore adopted herein. The moment distribution

expressed by Eq. (C.12) is slightly modified to

(C.14)

where , and the local abscissa of the section having is given by the

solution of

(C.15)

which has to be obtained separately according to the degree of the involved polynomials.

V s2( ) VE p L2--- s2–⎝ ⎠⎛ ⎞ P 1 h2–( )++=

M s2( ) ME2VEs2– p L s2–( )

s22---- P 1 h2–( )s2+–=

E1H E2H

V si( ) VE ξi p L2--- si–⎝ ⎠⎛ ⎞ P 1 hi–( )+–=

M si( ) MEiξiVEsi p L si–( )

si2--- P 1 hi–( )si+–+=

i 1 2,=( )

ξi

ξi11–⎩

⎨⎧

=forfor

i 1=i 2=

or ξi sid xd⁄=

M*

si M*

M si( ) MEiξiVEi

′ si p L si–( )si2---–+=

VEi

′ VE ξiP 1 hi–( )–= si* M*

M* M si*( ) MEi

ξiVEi

′ si* p L si

*–( )si

*

2----–+= =

Page 465: Seismic Assessment of Reinforced Concrete Frame Structures

442 Appendix C

Thus, for

(C.16)

which holds for non-uniform moments along , i.e. for .

For , the existence of solutions is controlled by the discriminant of the 2nd order Eq.

(C.15), which can be expressed by

(C.17)

If no real solutions exist, whereas if one or two solutions exist, given by

(C.18)

Once is obtained, it must be checked for its existance in the relevant interval ( ) in order

to be accepted as a possible solution and associated with the specific moving section under

study (cracking, yielding or null moment one). It may be transformed to the element abscissa,

according to or , respectively, for or . The subscript i has

been included inside brackets in order to identify the element part to which the moving

section is related. Actually, stands for the specific notation of each moving section, i.e. for

instance or as defined in 3.4.

Finally, it is worth mentioning that the case of uniform moments along one or both element

part(s) is not adressed because the concept of moving sections becomes meaningless once the

behaviour is fully controlled by end sections under the criteria and assumptions stated in 3.5.

Derivatives

The derivatives of moving section abscissas with respect to increments of end section

moments are detailed next, using the local abscissas as defined by Eqs. (C.16) or (C.18).

Since , it follows that , which renders

easier the task of obtaining derivatives because abscissas are written in terms of .

p 0=

si* M* MEi

ξiVEi

′----------------------=

EiH VEi

′ 0≠

p 0≠ ∆i( )

∆iL2--- ξi

VEi

p-------–

⎝ ⎠⎜ ⎟⎛ ⎞

2

2M* MEi

–( )

p---------------------------+=

∆i 0< ∆i 0≥

si* L

2--- ξi

VEi

p-------–

⎝ ⎠⎜ ⎟⎛ ⎞

∆i±=

si* EiH

x i( )* s1

*= x i( )* L s2

*–= i 1= i 2=

EiH

i( )

Ci κ+ Yi

x i( )*

si*

MEjMEj

0 ∆MEj+= ∂…( ) ∂∆MEj

( )⁄ ∂…( ) ∂MEj( )⁄=

MEj

Page 466: Seismic Assessment of Reinforced Concrete Frame Structures

Internal Force Distributions 443

Considering the relation between and , as well as the above introduced definitions of

and , and using the Krönecker symbol the following relations hold

(C.19)

(C.20)

As for the abscissas, situations of and are treated separately. For

(C.21)

(C.22)

which hold only if .

For , and in case of two existing real solutions , the general expressions for

derivatives become

(C.23)

(C.24)

whereas for the case of one double solution , Eq. (C.18) reduces to and

the respective derivatives are

(C.25)

x i( )* si

* VE

VEi

′ δij

x i( ) j,*

MEj∂∂x i( )

* MEj∂∂si

*

⎝ ⎠⎜ ⎟⎛ ⎞

si j,*= for i 1=

MEj∂∂si

*

⎝ ⎠⎜ ⎟⎛ ⎞

– si j,*–= for i 2=

⎩ ⎭⎪ ⎪⎪ ⎪⎪ ⎪⎨ ⎬⎪ ⎪⎪ ⎪⎪ ⎪⎧ ⎫

ξisi j,*= = =

MEj∂

∂VEi

′ξjL----⎝ ⎠⎛ ⎞–= and

MEj∂

∂MEi δij=

p 0= p 0≠ p 0=

si j,* ξj M* MEi

–( ) δijVEi

′ L–

ξiVEi

′ 2L

--------------------------------------------------------=

x i( ) j,* ξisi j,

* ξj M* MEi–( ) δijVEi

′ L–

VEi

′ 2L

--------------------------------------------------------= =

VEi

′ 0≠

p 0≠ ∆i 0>( )

si j,* ξiξj

pL--------- 1

p--- 1

2--- ξi

VEi

pL-------–

⎝ ⎠⎜ ⎟⎛ ⎞

ξiξj δij– ∆i-1/2±=

x i( ) j,* ξisi j,

* ξjpL------ 1

p--- 1

2--- ξi

VEi

pL-------–

⎝ ⎠⎜ ⎟⎛ ⎞

ξj ξiδij– ∆i-1/2±= =

∆i 0=( ) si* L

2--- ξi

VEi

p-------–

⎝ ⎠⎜ ⎟⎛ ⎞

=

si j,* ξiξj

pL---------= and x i( ) j,

* ξisi j,* ξj

pL------= =

Page 467: Seismic Assessment of Reinforced Concrete Frame Structures

444 Appendix C

Page 468: Seismic Assessment of Reinforced Concrete Frame Structures

Appendix D

The Event-to-Event Technique

In the present appendix, the event-to-event technique (Simons and Powell (1982), Porter and

Powell (1971)) is briefly introduced, following the steps for its application in the context of

classical stiffness based structural analysis.

A very simple example is used to illustrate the process. It consists of the single frame shown in

Figure D.1-a) where the non-linear behaviour is assumed concentrated in the six indicated

plastic hinges. Each of these hinges is ruled by a piecewise non-linear moment-rotation dia-

gram, as exemplified in Figure D.1-b) for the generic hinge i.

Figure D.1 The event-to-event scheme for stiffness based problems

At the local hinge level, the initial state point is assumed as ; upon application of the rota-

tion increment , the state point moves to . In between, events (stiffness changes) may

occur (for the illustrated example two events are considered, associated with moments and

), but only the first one is of interest. Thus, the bending moment and flexibility ,

associated with the first event at each hinge, are registered and the corresponding load incre-

ment reduction factor is set up as follows:

2

3

1

4

5 6 ∆F

∆Mi

∆θi

M1V

Mi0

M2V

θi0

V1

V2

θ

M

Hi0

Hi

Events: j=1,2

Mi* first Mj

V( )[ ]hinge i=j =1,2

fi*= flexib. after Mi

*

f1

f2

a) Structure b) Hinge i

Hi0

∆θi Hi

M1V

M2V Mi

*( ) fi*( )

Page 469: Seismic Assessment of Reinforced Concrete Frame Structures

446 Appendix D

(D.1)

The global reduction factor (r) is searched among all the hinges and given by

(D.2)

and the hinge number (possibly more than one) where the factor r occurs is denoted by .

If the load increment is reduced to and applied to the structure, this ensures

that non-linearity is about to occur only at the hinge(s) , whereas the remaining ones behave

with the current stiffness without reaching any event. The internal force and deformation state

of each section is updated accordingly, the flexibility of hinge(s) being modified to .

The new stiffness matrix can also be calculated and the remaining load increment, given by

(D.3)

is applied, following the same process as for the original increment .

Attention is drawn to the fact that no residual forces are likely to appear in this process,

because the subdivision is enforced exactly to avoid them, i.e. to closely follow the equilibrium

path.

However, it is clear that, such apparent “advantage” may easily become a major drawback

when a significant number of plastic hinges has entered non-linear behaviour, because they all

contribute directly to the global equilibrium equations. Different constitutive relationships and

non-concomitant events among all sections obviously induce a heavy subdivision process

which is more pronounced as the number of non-linear sections increases. Mostly at critical

loading stages (like reversals or zero crossings), any small out-of-phase loading between dif-

ferent sections may enforce many small steps that, in a usual N-R process, would not be

needed. Therefore, the use of the event-to-event scheme, in the context of non-linear global

structure analysis, has to be judged carefully in view of the local constitutive relationship and,

mostly, of the number of sections involved.

riMi

* Mi0–

∆Mi--------------------=

r min ri{ } i=1,...,Nhinge( )=

ir

F∆ F∆ r r F∆=

ir

ir fir*( )

F∆ c F∆ F∆ r–=

F∆

Page 470: Seismic Assessment of Reinforced Concrete Frame Structures

Appendix E

Non-Linear Dynamic and Static Analysis Scheme

The analysis of both static and dynamic non-linear structural problems can be performed by

means of a common scheme, since the dynamic one is transformable into a pseudo-static prob-

lem (in each load step) if the dynamic equilibrium equations are integrated step-by-step using

well established algorithms. Hence, in this appendix, the non-linear dynamic scheme is

recalled, in the context of the Newmark method family, currently one of the most widely

accepted integration techniques, and then particularized to both the static case and the seismic

one, the latter being a specific dynamic loading type.

For notation clarity in the following, the global structure displacement vector will be sim-

ply denoted by and the general (static or dynamic) applied nodal forces by . Thus, in gen-

eral and , and the semi-discrete dynamic equilibrium equation system is

given by

(E.1)

where and stand for the global structure mass and damping matrices, respectively, and

is the restoring force vector associated with the displacement configuration ; for the

elastic case is simply given by , where refers here to the global elastic

stiffness matrix.

Let the time instants and correspond to steps and , respectively, i.e.

and , where is the time step interval, such that a common notation

holds valid for both dynamic and static analysis. The discretization of the continuous time var-

iable into finite steps allows to write Eq. (E.1) in the discrete form

uG

d q

d d t( )= q q t( )=

M d··⋅ C d·⋅ r d( )+ + q=

M C

r d( ) d

r d( ) r K0 d⋅= K0

t ∆t– t k 1– k

t ∆t– tk 1–= t tk= ∆t

t

Page 471: Seismic Assessment of Reinforced Concrete Frame Structures

448 Appendix E

(E.2)

where the following approximations were introduced for step

(E.3)

Step-by-step methods for the integration of Eq. (E.2) typically aim at the solution for step ,

which can be obtained either exclusively in terms of the response for the previous step ,

as in the case of explicit methods, or also in terms of the response for step , leading to

implicit algorithms.

Explicit methods are computationally more attractive but their application is often limited by

response stability conditions, which may require rather small particularly for structures

with significant contributions from high frequency vibration modes. By contrast, implicit

methods have the major advantage of unconditional stability, for which the time step is exclu-

sively chosen for accuracy purposes, but more elaborate algorithms are required.

The most common implicit methods are essentially based on the classical Newmark method

(Delgado (1984)), where an average constant acceleration is assumed in the interval

and is defined by

(E.4)

which, upon integration leads to

(E.5)

(E.6)

Parameters and are introduced to control the accuracy and stability of the method.

M ak⋅ C vk⋅ rk+ + qk=

k

dk d tk( )≈ d t( )=

vk d· tk( )≈ d· t( )=

ak d·· tk( )≈ d·· t( )=

rk r d tk( )( )≈ r d t( )( )=

qk q tk( ) q t( )= =

k

k 1–( )

k

∆t

tk 1– tk,[ ]

ak 1 γ–( )ak 1– γak+=

vk vk 1– 1 γ–( )ak 1– γak+[ ]∆t+=

dk dk 1– vk 1– ∆t 12--- β–⎝ ⎠⎛ ⎞ ak 1– βak+ ∆t2+ +=

β γ

Page 472: Seismic Assessment of Reinforced Concrete Frame Structures

Non-Linear Dynamic and Static Analysis Scheme 449

The particular case of and leads to the well known central difference method,

an explicit and conditionally stable one.

For the method is unconditionally stable, although the desirable second order

accuracy (i.e. second order derivatives accurately reproduced) is only achieved for .

On the other hand, numerical damping of contributions due to high frequency spurious modes

can be obtained if , but only first order accuracy can be obtained. Therefore, numerical

dissipation and second order accuracy cannot be achieved simultaneously, which is a drawback

that has motivated improvements in the method over the last two decades (Hilber et al. (1977))

in order to introduce numerical dissipation for higher spurious modes, while keeping good

response performance at the lower frequency ones. Further references to such improvements

are also included in 5.2 in the context of the experimental pseudo-dynamic method, but in the

following paragraphs only the classical version of the Newmark method is considered aiming

at a simpler presentation of the non-linear dynamic analysis scheme.

Nevertheless, Hughes (1987) has shown that the best efficiency in numerical dissipation is

obtained by adopting , for a given , which, for the case where sec-

ond order accuracy is to be ensured, gives and , indeed the most common

values in the Newmark method parameters.

Eqs. (E.5) and (E.6) can be rearranged in terms of the displacement increment

and the kinematic entities of the previous step as follows

(E.7)

(E.8)

where and depend exclusively on the response for the step and are given,

respectively, by

(E.9)

β 0= γ 1 2⁄=

2β γ 1 2⁄≥ ≥

γ 1 2⁄=

γ 1 2⁄>

β γ 1 2⁄+( )2 4⁄= γ 1 2⁄≥

γ 1 2⁄= β 1 4⁄=

∆dk dk dk 1––=

ak1

β∆t2-----------∆dk ak 1––=

vkγ

β∆t---------∆dk vk 1––=

ak 1– vk 1– k 1–( )

ak 1–1β∆t---------vk 1–

12β------ 1–⎝ ⎠⎛ ⎞ ak 1–+=

Page 473: Seismic Assessment of Reinforced Concrete Frame Structures

450 Appendix E

(E.10)

Introducing Eqs. (E.9) and (E.10) in Eq. (E.2), the following equilibrium equation is obtained

(E.11)

with

(E.12)

Since the restoring force vector depends non-linearly of , Eq. (E.11) must be

solved by means of an iterative process such that, when convergence is reached, the following

residual force vector vanishes to within a pre-defined tolerance

(E.13)

Thus, let be the approximation of for the iteration , for which the corre-

sponding force residuals are . Developing in Taylor series for

and denoting by the displacement variation such that

(E.14)

the following first order approximation can be obtained for the residuals corresponding to iter-

ation

(E.15)

Hence, for the linearization expressed by Eq. (E.15), the force residuals vanish if the displace-

ment variation vector is such that

(E.16)

vk 1–γβ--- 1–⎝ ⎠⎛ ⎞ vk 1–

γ2β------ 1–⎝ ⎠⎛ ⎞ ∆tak 1–+=

1β∆t2-----------M γ

β∆t---------C+⎝ ⎠

⎛ ⎞ ∆dk⋅ r dk 1– ∆dk+( )+ qk=

qk qk M ak 1–⋅ C vk 1–⋅+ +=

r d t( )( ) d t( )

Ψ ∆dk( )

Ψ Ψ ∆dk( ) qk r dk 1– ∆dk+( ) 1β∆t2-----------M γ

β∆t---------C+⎝ ⎠

⎛ ⎞ ∆dk⋅––= =

∆dkn 1– ∆dk n 1–

Ψkn 1– Ψ ∆dk

n 1–( )= Ψ ∆dkn 1–

δkn

∆dkn ∆dk

n 1– δkn+=

n

Ψ ∆dkn( ) Ψ ∆dk

n 1–( ) ∂Ψ∂ ∆dk( )-----------------

n 1–δk

n⋅+=

δkn

∂Ψ∂ ∆dk( )-----------------

n 1–– δk

n⋅ Ψ ∆dkn 1–( )=

Page 474: Seismic Assessment of Reinforced Concrete Frame Structures

Non-Linear Dynamic and Static Analysis Scheme 451

Taking into account that , the above expressed derivatives can

be obtained from Eq. (E.13) and, therefore, Eq. (E.16) transforms into

(E.17)

where the so-called effective stiffness matrix is given by

(E.18)

in which is the tangent stiffness matrix for the iteration of step .

Thus, according to Eq. (E.13) the residual vector is evaluated by

(E.19)

where and the displacement correction becomes

(E.20)

which can be replaced in Eq. (E.14) to provide an estimate for iteration . The process contin-

ues until the satisfaction of a given convergence criterion, which can be defined in terms of a

norm or the maximum of residual forces, or in terms of a residual displacement norm or even

in terms of residual energy (Owen and Hinton (1980), Faria (1994)).

The above stated sequence of steps constitutes the Newton-Raphson method, widely used to

solve non-linear equation systems, for which an initial estimate must be provided. As

referred in Faria (1994) this estimate is somewhat arbitrary since the subsequent corrections

(given by Eq. (E.20)) will tend to approach the real solution. Obviously, the more reasonable

the estimate, the less iterations will be needed to reach convergence; in this context, the option

for either or appear to be adequate, and the first one is in fact adopted

in CASTEM2000. Thus, replacing by in Eq. (E.19), the first estimate of residu-

als is given by

∂ …( ) ∂ ∆dk( )⁄ ∂ …( ) ∂d⁄[ ]k=

Kkn 1– δk

n⋅ Ψkn 1–=

Kkn 1–

Kkn 1– 1

β∆t2-----------M γ

β∆t---------C d∂

∂r⎝ ⎠⎛ ⎞

k

n 1–

+ +=

∂r ∂d⁄( )kn 1– n 1– k

Ψkn 1–

Ψkn 1– qk rk

n 1– 1β∆t2-----------M γ

β∆t---------C+⎝ ⎠

⎛ ⎞ ∆dkn 1–⋅––=

rkn 1– r dk 1– ∆dk

n 1–+( )=

δkn Kk

n 1–[ ]1–Ψ⋅ k

n 1–=

n

∆dk0

∆dk0 0= ∆dk ∆dk 1–=

∆dk ∆dk0 0=

Page 475: Seismic Assessment of Reinforced Concrete Frame Structures

452 Appendix E

(E.21)

which then provides the first displacement correction using Eq. (E.20).

The tangent stiffness matrix included in is often replaced by a stiffness matrix which is

kept constant during several steps or iterations in order to reduce the computational effort

inherent in the stiffness matrix factorization; this leads to the so-called modified Newton-

Raphson schemes which can also be easily used in CASTEM2000.

The analysis for static loading is, indeed, a particular case of the dynamic problem: since accel-

eration and velocity vectors vanish, the residual vector as given by Eq. (E.19) for the generic

iteration becomes and the matrix coincides with the tangent

stiffness matrix , while displacement corrections are still obtained by Eq. (E.20).

Additionally, for the first iteration with the force residuals match the increment of

external applied forces, i.e., , because the convergence

in the previous step imposed the satisfaction of to within a pre-defined toler-

ance (which justifies the use of in the previous expression of ).

The seismic input is a particular case of dynamic loading in which ground motions are imposed

to the structure base supports (or foundation). It can be derived directly from the D´Alembert

principle (Clough and Penzien (1975)) to yield the most common way of considering the seis-

mic input. Thus, denoting by the vector of the base translational displacement components,

the total displacement , velocity and acceleration vectors of the structure

become

(E.22)

where is the so-called pseudo-static matrix reflecting the rigid body modes due to ground

motion.

Re-writing Eq. (E.1) in terms of total kinematic entities and assuming that no other forces

are applied, it yields

Ψk0 qk r dk 1–( )–=

Kkn 1–

n 1– Ψkn 1– qk rk

n 1––= Kkn 1–( )

∂r ∂d⁄( )kn 1–

∆dk0 0=

Ψk0 qk r dk 1–( )– qk qk 1––≈ ∆qk= =

r dk 1–( ) qk 1–=

≈ Ψk0

dg

dT( ) d· T( ) d··T( )

dT d 1˜

dg⋅+=

d· T d· 1˜

d· g⋅+=

d··T d·· 1˜

d··g⋅+=

q

Page 476: Seismic Assessment of Reinforced Concrete Frame Structures

Non-Linear Dynamic and Static Analysis Scheme 453

(E.23)

where the contribution of restoring forces vanishes because is strictly associ-

ated with rigid body motions of the structure.

Additionally, for most cases of building structure analysis, the damping matrix is defined in

terms of the stiffness one and the contribution of also vanishes. Thus, Eq. (E.23) is

currently used in the form

(E.24)

which is equivalent to Eq. (E.1) with the dynamic force vector particularized for

. The pseudo-static matrix is the well known matrix consisting of unit val-

ues for the displacement components parallel to each of the translational components of

and zero in the remaining positions.

In the present work, this form of seismic action prescription has been adopted, even in cases

when is also defined in terms of . Indeed, although not strictly valid in such circum-

stances, this option has nevertheless been kept based on the assumption that damping forces

are essentially caused by relative motions and on the belief that, for non-linear behaviour of

reinforced concrete structures, the damping source arises mainly from hysteresis (supposed

already included in ) thus allowing to eliminate contributions involving the viscous

damping matrix .

M d··⋅ C d·⋅ r d( )+ + M– 1˜

d··g⋅ ⋅ C 1˜

d· g⋅ ⋅– r 1˜

dg⋅( )–=

r 1˜

dg⋅( ) 1˜

dg⋅

C

C 1˜

d· g⋅ ⋅

M d··⋅ C d·⋅ r d( )+ + M– 1˜

d··g⋅ ⋅=

q M– 1˜

d··g⋅ ⋅= 1˜

d··g

C M

r d( )

C

Page 477: Seismic Assessment of Reinforced Concrete Frame Structures

454 Appendix E