You are on page 1of 282

To Professor Jacques Heyman

Abstract
This dissertation proposes a novel formulation for the rocking motion (RM) of rigid-blocks
when no sliding mechanisms are present. Both theoretical and numerical analyses are included.
The results obtained from a new formulation for single block dynamics, which have shown to
be extremely powerful, led to the generalization for many degrees of freedom systems presented
herein. That generalization was possible through a transformation of the Euclidean congura-
tion space to a complex Riemannian manifold endowed with an Hermitian metric tensor.
The present contribution applies also techniques derived from chaos theory to the RM problem
for the understanding and quantication of the dynamic stability. The probability of overturn of
rocking blocks under random loading is both experimentally and numerically investigated.
Through an intensive use of the Poincar surface of section technique and bifurcation analysis,
the underlying structure of the phase space is highlighted. Delay coordinated, recurrence plots
and recurrence quantication analysis are used to nd the recurrences and patterns present in
the dynamics at dierent time scales. An estimator for the quantication of chaos in the prob-
lem of rocking blocks under earthquake loading is proposed.
Stochastic analysis is performed through an ensemble of input earthquake samples. The ensem-
ble limits and other stochastic properties of the relevant physical magnitudes are analyzed.
A set of experiments at a seismic table on four blue granite stones were conducted to validate
the theoretical analyses. The response is investigated under dierent input actions and block
geometries. Several applications to Earthquake Engineering and structural safety assessment
are therefore derived. In particular, numerical thresholds for the probability of collapse of slen-
der structures are found by means of dierent criteria. A model for the probability of collapse
based on the stationary solution of the associated Fokker-Planck equation is proposed.
iii
Acknowledgment
This work was partially supported by the FCT (Fundao para a Cincia e a Tecnologia)
through Grant with Ref. SFRH/BD/9014/2002.
I would like to express my acknowledgment to my supervisors in Portugal and Spain as well as
to the personnel of LNEC (Laboratrio Nacional de Engenharia Civil), Portugal, for their kind
support and advice during this work, specially Dr. Alfredo Campos Costa.
I am also in debt with Professor Pere Roca from UPC (Universitat Politecnica de Catalunya),
Spain.
iv
Contents
I Introduction and Critical Review of Classical Theory 1
1 Introduction 2
1.1 Rocking Motion and Historical Constructions . . . . . . . . . . . . . . . . . . 2
1.2 Models for Masonry Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 An Overview of the Existing Works related to RM . . . . . . . . . . . . . . . . 12
1.4 The Present Contribution. Objectives and Scope . . . . . . . . . . . . . . . . 14
2 Critical Review of the Classical Theory 16
2.1 PWT Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2 PWT Model for Impact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
II Formulation for Multiple Degree of Freedom Systems 23
3 Geometrization of RM Multi-Block Dynamics 24
3.1 Outline of the Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2 Set-up of the Conguration Space . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3 Implementation of Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.4 Extension to the Riemannian Conguration Space . . . . . . . . . . . . . . . . 35
3.5 General Procedure to Obtain the Reduced Equations of Motion . . . . . . . . . 39
3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
v
4 Formulation for Planar Systems 44
4.1 General Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.2 Single Degree of Freedom Systems . . . . . . . . . . . . . . . . . . . . . . . . 45
4.3 Multiple Degree of Freedom Systems . . . . . . . . . . . . . . . . . . . . . . 49
4.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
III Single Block Dynamics and Stability 54
5 Complex Formulation for Single Block Systems 55
5.1 The RM limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.2 Free Rocking Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.3 Implementation of Impact and Seismic Actions . . . . . . . . . . . . . . . . . 59
5.4 Mechanical Analogies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.5 Comparison with Housner Theory . . . . . . . . . . . . . . . . . . . . . . . . 67
5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6 Analysis of the Response 71
6.1 Description of the Experimental Research . . . . . . . . . . . . . . . . . . . . 71
6.2 Types of Response and Their Spectra . . . . . . . . . . . . . . . . . . . . . . . 77
6.3 Reproducibility of the Response . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.4 Classication of Orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
7 Stability Analysis 99
7.1 Maximum Angles and Overturn . . . . . . . . . . . . . . . . . . . . . . . . . 100
7.2 Ensemble Average Values of Velocity and Energy . . . . . . . . . . . . . . . . 104
7.3 Correlation Maps for Experimental Maximum Angles . . . . . . . . . . . . . . 111
7.4 Number of Impacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
7.5 Evaluation of First Passage Time . . . . . . . . . . . . . . . . . . . . . . . . . 128
7.6 Stability in the Amplitude-Frequency Range . . . . . . . . . . . . . . . . . . . 128
vi
7.7 Bifurcation Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
7.8 Delay-coordinates Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
7.9 Poincar Surface of Section Analysis . . . . . . . . . . . . . . . . . . . . . . . 144
7.10 PSS for Harmonic Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
7.11 Eect of White Noise on the PSS . . . . . . . . . . . . . . . . . . . . . . . . . 160
7.12 PSS for Random Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
7.13 Recurrence Plots and Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . 169
7.14 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
IV Applications to Earthquake Engineering 196
8 Probability of Collapse for SDOF Structures 197
8.1 Introduction: The Rocking Spectrum . . . . . . . . . . . . . . . . . . . . . . . 197
8.2 Experimental Collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
8.3 Probability Curves for Overturn . . . . . . . . . . . . . . . . . . . . . . . . . 202
8.4 Analytical Model for the Probability of Collapse . . . . . . . . . . . . . . . . . 205
8.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
V Conclusions and Future Research 216
9.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
9.2 Future Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
VI Annexes 221
A Periods of a Curved-Base Block 222
B Mathematical Results and Formulae 225
C Dirac-Delta Forces and Phase Dynamics 229
vii
D Derivation of the associated Fokker-Planck Equation 233
E Experimental Details 237
F Sources of Error and Their Eects 242
G Computer Programs 251
viii
ABBREVIATIONS
The following abbreviations are used in this work.
CF Complex Formulation
DEM Discrete Element Method
DOF Degree Of Freedom
FEM Finite Element Method
FPE Fokker-Plank Equation
FPT First Passage Time from the static potential well
MDOF Multiple Degree Of Freedom
ODE Ordinary Dierential Equation
PC Probability of Collapse
PGA Peak Ground Acceleration
PSS Poincar Surface of Section method
PWT Piecewise Theory
RM Rocking Motion
RP Recurrence Plot
RQA Recurrence Quantication Analysis
SDOF Single Degree Of Freedom
UPO Unstable Periodic Orbit
ix
NOTATION
The notation used in this work is fairly standard, with two important exceptions.
1. It is customary to use column matrices for vectors and row-type matrices for bases. In
this work, however, the opposite convention is assumed. The reason for this change
in notation is the economy of space and notation since along the computations, vectors
appear much more frequently than bases. Obviously, both formulations are equivalent.
2. The symbol t is used for the non-dimensional time (to be introduced in Chapter 2) as
it appears much more frequently than the usual dimensional time. For the dimensional
time, the symbol is assigned.
Einstein Summation Convention is followed through the whole work unless it is clearly stated
that there is no sum over repeated indexes.
General denitions
Rotation angle of rocking block
Time
r || Complex modulus of
Complex argument of
Single block geometry
b Half-width of the block
h Half-height of the block
h/b Slenderness ratio
R Distance from center of rotation to center of mass
M Mass of the block
I Moment of inertia measured from O or O

x / Normalized angle
|x| Non-dimensional complex modulus of x
x
Multi-block geometry
x Position of a particle in the body reference-frame
< u, v > Inner product of two vectors

d
ii+1
Position of a hinge of body i + 1 in the reference-frame of body i
tr(A) Trace of matrix A
A
T
Transpose of matrix A
det(A) Determinant of matrix A
v(n) Number of arista (possible mechanisms) of body n
n Mechanism index of body n
Static friction angle
Rocking Motion parameters
Static critical angle, given by; = arctan(b/h)
p Frequency parameter
Angular velocity reduction ratio
Stability Analysis
d Time lag for delay coordinates
W
s
Stable manifold
W
u
Unstable manifold
External action
a() Horizontal acceleration
a
g
a()/g Normalized horizontal acceleration
a
w
g tan() West acceleration
Factor for horizontal excitation amplitude under harmonic motion
External frequency of horizontal excitation
/p Normalized frequency of horizontal excitation
T
H
Period using Housner formula
f
H
Housner frequency f
H
= 1/T
H
xi
Derived functions
T Kinetic function
U Potential function
L Lagrangian function
R Routhian
L Non-dimensional Lagrangian function
H Hamiltonian
F Generating function in point canonical transformation
S Generating function in Hamilton-Jacobi theory
E Mechanical energy
e E/I p
2
Non-dimensional mechanical energy
k T/I p
2
Non-dimensional kinetic energy
Q
k
Generalized force corresponding to coordinate "k"

j
New constant canonical coordinates in Hamilton-Jacobi Theory
Mathematical functions

nm
=
nm
=
m
n
Kronecker delta symbol

n
(x) Delta-sequence
(x) Dirac-delta function
(x) Heaviside step function
Probability measures
< > Expected value operator over the sample space
F(t) Mean value of F(t) in the time domain
< T
e
> Mean rst passage time
N() Number of impacts during time
N

= N()/ Number of impacts per time unit


n
p
Stationary limit of N

Constants
g Acceleration of gravity
RM parameter. Conserved angular momenta in the complex plane
xii
List of Figures
1.1 Ruins of Convento do Carmo at the city Lisbon, Portugal. . . . . . . . . . . . . 3
1.2 Iberian Peninsula hazard map . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 The masonry material. Types and congurations. . . . . . . . . . . . . . . . . 6
1.4 Block with a curvature at its base . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.5 RM model for single block dynamics . . . . . . . . . . . . . . . . . . . . . . . 11
1.6 Modal analysis model for a single block . . . . . . . . . . . . . . . . . . . . . 11
1.7 Discrete-element method for single block dynamics . . . . . . . . . . . . . . . 11
2.1 Geometry and notation for the rocking block. . . . . . . . . . . . . . . . . . . 17
2.2 Energetic map for a RM block . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.1 Reference frames for rigid rotations of two-block assemblies . . . . . . . . . . 26
3.2 Rotation about three fracture lines. . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3 Dynamical symmetry associated to the sign of . . . . . . . . . . . . . . . . . 34
3.4 Unsymmetric non-prysmatic rocking Block . . . . . . . . . . . . . . . . . . . 34
3.5 Single block conguration space: r = ||, = arg . . . . . . . . . . . . . . . 36
3.6 Procedure for the determination of the Probability of collapse of masonry struc-
tures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.7 Masonry wall under out-of-plane loading. Collapse mechanism . . . . . . . . . 41
4.1 Closed-kinematic chain structure. Denition of parameters . . . . . . . . . . . 46
4.2 Closed-kinematic chain structure. Solution after DOF reduction . . . . . . . . 47
xiii
4.3 Line of thrust of a masonry arch under lateral load proportional to its own
weight. Solution computed with ARCOTSAM.mws . . . . . . . . . . . . . . . 48
4.4 Collapse mechanism I of a four-hinged masonry arch. Solution after DOF re-
duction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.5 Collapse mechanism II of a four-hinged masonry arch . . . . . . . . . . . . . . 49
4.6 Collapse mechanism of two rigid stacked blocks . . . . . . . . . . . . . . . . . 51
4.7 Column formed by 14 equal stacked prismatic blocks . . . . . . . . . . . . . . 52
4.8 Planar motion of of a column formed by 14 stacked prismatic blocks . . . . . . 53
5.1 Impulsive forces during impact . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.2 Potential for an inverted pendulum system . . . . . . . . . . . . . . . . . . . . 66
5.3 Eective potential for two-body problem . . . . . . . . . . . . . . . . . . . . . 67
5.4 Eect of the rocking limit on the response . . . . . . . . . . . . . . . . . . . . 69
5.5 Comparison between complex formulation and Housner theory . . . . . . . . . 69
6.1 Experimental energies for sine-sweep tests . . . . . . . . . . . . . . . . . . . . 72
6.2 Specimens for experimental tests. . . . . . . . . . . . . . . . . . . . . . . . . 74
6.3 Geometry of the testing specimens . . . . . . . . . . . . . . . . . . . . . . . . 75
6.4 Fourier spectrum of free RM. Specimen 1 . . . . . . . . . . . . . . . . . . . . 77
6.5 Fourier spectrum of free RM. Specimen 2 . . . . . . . . . . . . . . . . . . . . 78
6.6 Comparison theory-experiment for free RM on specimens 1 and 2 with tted
parameters of Table E.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.7 Comparison theory-experiment for free RM on specimens 3 and 4 with tted
parameters of table E.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.8 Transient and stationary parts of experimental response under constant ampli-
tude harmonic load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
6.9 Stationary part of experimental periodic modes . . . . . . . . . . . . . . . . . 81
6.10 Fourier spectrum of harmonic loading with constant amplitude . . . . . . . . . 81
6.11 Comparison theory-experiment for constant amplitude harmonic forcing on spec-
imen 1 with tted parameters of table E.1 . . . . . . . . . . . . . . . . . . . . 82
xiv
6.12 Spectrumof the block response for a hanning sinusoidal load with amplitude=6mm
and dierent frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.13 Comparison theory-experiment for hanning sine on specimen 1 with tted pa-
rameters of table E.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.14 Typical input series for a random load test . . . . . . . . . . . . . . . . . . . . 85
6.15 Ensemble of sample input functions . . . . . . . . . . . . . . . . . . . . . . . 86
6.16 Ensemble mean of a
g
= a/g as a function of time for two ensembles of 100 and
275 samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.17 Histograms of the ensemble at dierent sampling times. . . . . . . . . . . . . . 88
6.18 Ensemble mean and variance of a
g
= a/g as function of time . . . . . . . . . . 89
6.19 Recurrence plot and autocorrelation function of sample input series . . . . . . . 89
6.20 Power spectral density of the sample input series . . . . . . . . . . . . . . . . 90
6.21 Spectrum of the response under random loading . . . . . . . . . . . . . . . . . 91
6.22 Comparison theory-experiment for random loading on specimen 3 I. . . . . . . 92
6.23 Comparison theory-experiment for random loading on specimen 3 II. . . . . . . 92
6.24 Repeatability of the experimental response for harmonic loading . . . . . . . . 93
6.25 Non experimental repeteability for Random motion . . . . . . . . . . . . . . . 93
6.26 Non experimental repeteability for Random motion II . . . . . . . . . . . . . . 93
6.27 Time series orbit and phase-space trajectory of periodic response, (1,1) Mode . 95
6.28 Periodic response. Time series orbit and phase portrait. (1,3) mode . . . . . . . 95
6.29 Periodic response. Time series orbit and phase portrait. (1,7) mode . . . . . . . 96
6.30 Quasi-periodic response. Time series orbit and phase portrait . . . . . . . . . . 96
6.31 Chaotic response. Time series orbit and phase portrait . . . . . . . . . . . . . . 97
6.32 Overturning response. Time series orbit and phase portrait . . . . . . . . . . . 98
7.1 Experimental maximum angle for harmonic load vs external amplitude and fre-
quency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
7.2 Maximum angle peaks for constant external amplitude. Numerical simulation. . 102
7.3 Experimental maximum angles (absolute value) of specimen 4 under random load103
xv
7.4 Experimental distribution of ensemble mean of || for specimen 3 under random
loading. Eect of PGA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.5 Schematic graph of the experimental stationary limit of < || > under random
loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
7.6 Experimental stationary limit of < || > for specimen 3 under random loading . 105
7.7 Experimental distribution of velocity sampling . . . . . . . . . . . . . . . . . . 105
7.8 Dependence of < v > with PGA. Experimental tests on specimen 3 . . . . . . . 106
7.9 Experimental potential, kinetic and mechanical energies of specimen 3 under
random loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
7.10 Experimental mechanical energy of specimen 3 under random loading for dif-
ferent PGA values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.11 Ensemble mean of energy for two values . . . . . . . . . . . . . . . . . . . . 108
7.12 Stationary limit of < e > vs p . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
7.13 Stationary limit of < e > vs . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.14 Stationary limit of < e > vs PGA . . . . . . . . . . . . . . . . . . . . . . . . . 111
7.15 Map of m
n
vs m
n+1
for Free RM. Experiment . . . . . . . . . . . . . . . . . . . 112
7.16 Map of m
n
vs m
n+1
for Harmonic load. Experiment . . . . . . . . . . . . . . . 114
7.17 Map of m
n
vs m
n+1
for Harmonic load. Experiment . . . . . . . . . . . . . . . 116
7.18 Map of correlation of consecutive maxima for random load under dierent PGA
values. Experimental test on specimen 3 . . . . . . . . . . . . . . . . . . . . . 117
7.19 Autocorrelation function of experimental maxima of ||. Random loading . . . 119
7.20 Experimental number of impacts per time for free rocking regime of SP 1 . . . 120
7.21 Experimental number of impacts as function of time. RUN-UP test . . . . . . . 121
7.22 Experimental number of impacts as function of time. RUN-DOWN test . . . . 122
7.23 Experimental number of impacts as function of time. Harmonic loading . . . . 122
7.24 Experimental number of impacts as function of time. Random loading . . . . . 123
7.25 Number of impacts as function of time. Random loading for numerical simula-
tion. Eect of PGA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
7.26 Stationary limit of the number of impacts per time as function of . . . . . . . 125
xvi
7.27 Stationary limit of the number of impacts per time as function of p . . . . . . . 126
7.28 Stationary limit of the number of impacts per time as function of . . . . . . . 126
7.29 Stationary limit of the number of impacts per time as function of external PGA 127
7.30 Ensemble average of First Passage Times for random loading. Numerical sim-
ulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
7.31 Numerical overturn thresholds for external amplitude and frequency in har-
monic loading regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
7.32 Stability amplitude-frequency maps. Eect of the slenderness ratio. . . . . . . 132
7.33 Stability amplitude-frequency maps. Eect of the frequency parameter. . . . . 133
7.34 Stability amplitude-frequency maps. Eect of damping. . . . . . . . . . . . . . 134
7.35 Stability amplitude-frequency maps. Eect of the external phase . . . . . . . . 136
7.36 Bifurcation. Amplitude scanning for dierent values and = 0.925, 1.0 . . . 138
7.37 Bifurcation. scanning for dierent values . . . . . . . . . . . . . . . . . . 139
7.38 Bifurcation plot for scanning with two values of amplitude, = 10.0 and
= 0.8, 0.925, 1.0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
7.39 Bifurcation plot for scanning. = 10.0 . . . . . . . . . . . . . . . . . . . . 140
7.40 Delay plots for periodic modes under harmonic forcing. Numerical simulation . 142
7.41 Delay plots of periodic (1,3) Mode for dierent time lags . . . . . . . . . . . . 143
7.42 Delay plots for quasi-periodic modes under harmonics forcing with constant
amplitude. Numerical simulation . . . . . . . . . . . . . . . . . . . . . . . . . 143
7.43 Delay plot for chaotic orbit under harmonic forcing with constant amplitude.
Numerical simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
7.44 Singular points in the Phase plane . . . . . . . . . . . . . . . . . . . . . . . . 146
7.45 Denition of Stroboscopic Poincare map . . . . . . . . . . . . . . . . . . . . . 147
7.46 Periodic modes in the reduced (X1, X2) state space . . . . . . . . . . . . . . . 149
7.47 Unstable periodic modes in the reduced (X1, X2) state space . . . . . . . . . . 150
7.48 Poincar section for quasi-periodic modes . . . . . . . . . . . . . . . . . . . . 151
7.49 PSS. Eect of damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
7.50 PSS. Eect of external amplitude for undamped systems I . . . . . . . . . . . . 153
xvii
7.51 PSS. Eect of external amplitude for undamped systems II . . . . . . . . . . . 154
7.52 PSS. Eect of external amplitude for damped motion I . . . . . . . . . . . . . 155
7.53 PSS. Eect of external amplitude for damped motion II . . . . . . . . . . . . . 155
7.54 PSS. Eect of external frequency for undamped motion I . . . . . . . . . . . . 156
7.55 PSS. Eect of external frequency for undamped motion II . . . . . . . . . . . . 156
7.56 PSS. Eect of external frequency for damped systems I . . . . . . . . . . . . . 157
7.57 PSS. Eect of external frequency for damped systems II . . . . . . . . . . . . 157
7.58 PSS. Eect of slenderness ratio for undamped systems . . . . . . . . . . . . . 158
7.59 PSS. Eect of slenderness ratio for damped systems . . . . . . . . . . . . . . . 159
7.60 Eect of external phase in the PSS. Undamped . . . . . . . . . . . . . . . . 159
7.61 Eect of external phase in the PSS. Damping: = 0.925 . . . . . . . . . . . 160
7.62 Typical input load of a white-noise analysis . . . . . . . . . . . . . . . . . . . 161
7.63 Eect of white noise on subharmonic (1,3) Mode . . . . . . . . . . . . . . . . 161
7.64 Eect of white noise on sinusoidal input load for a chaotic orbit . . . . . . . . 162
7.65 Eect of moderate noise level on the stroboscopic PSS for damped block. . . . 163
7.66 Denition of separatrix for PSS under random loading . . . . . . . . . . . . . 163
7.67 PSS for random loading. Eect of damping . . . . . . . . . . . . . . . . . . . 164
7.68 PSS for random loading. Eect of PGA . . . . . . . . . . . . . . . . . . . . . 165
7.69 PSS for random loading. Eect of slenderness . . . . . . . . . . . . . . . . . . 166
7.70 Two types of stroboscopic PSS for random loading . . . . . . . . . . . . . . . 167
7.71 Number of Poincar points crossing the separatrix as function of the external
PGA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
7.72 Analysis of attractor structure. Main regions . . . . . . . . . . . . . . . . . . . 168
7.73 Probability distribution function for random loading in the phase space . . . . . 169
7.74 Recurrence plot of () under random loading for four dierent embedding di-
mensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
7.75 Recurrence plot of () under random loading for 4 dierent delay times . . . . 171
7.76 Recurrence plot of stable periodic (1,1) Mode . . . . . . . . . . . . . . . . . . 172
7.77 Recurrence plots of subharmonic (1,3) Mode . . . . . . . . . . . . . . . . . . 173
xviii
7.78 Main distance in the RP graph of subharmonic (1,3) Mode . . . . . . . . . . . 173
7.79 Main distance in the RP graph of subharmonic (1,5) Mode . . . . . . . . . . . 174
7.80 Recurrence plot of unstable subharmonic (3.9) Mode . . . . . . . . . . . . . . 174
7.81 Basic structure in the recurrence plot of unstable subharmonic (3,9) Mode . . . 175
7.82 Recurrence plot of 3-island quasi-periodic mode . . . . . . . . . . . . . . . . . 175
7.83 Recurrence plots of quasi-periodic modes. 5 and 7 island structures. . . . . . . 176
7.84 Basic pattern of the quasi-periodic 3-island . . . . . . . . . . . . . . . . . . . 177
7.85 Basic pattern of the quasi-periodic 7-island . . . . . . . . . . . . . . . . . . . 177
7.86 Recurrence plot of a chaotic orbit under harmonic loading. . . . . . . . . . . . 178
7.87 Basic structure of the RP of a chaotic orbit under harmonic forcing. . . . . . . 178
7.88 Recurrence plots for random loading for two PGA values. . . . . . . . . . . . . 180
7.89 Recurrence plots for random loading of a typical collapse orbit . . . . . . . . . 181
7.90 Mean RR versus external PGA in the recurrence plot . . . . . . . . . . . . . . 185
7.91 Mean DET versus external PGA in the recurrence plot . . . . . . . . . . . . . 186
7.92 Mean L versus external PGA in the recurrence plot . . . . . . . . . . . . . . . 186
7.93 Mean DIV versus external PGA in the recurrence plot . . . . . . . . . . . . . . 187
7.94 Mean LAM versus external PGA in the recurrence plot . . . . . . . . . . . . . 188
7.95 Mean TT versus external PGA in the recurrence plot. Eect of PGA . . . . . . 188
7.96 Ensemble means of the L
entr
and V
entr
estimators versus external PGA in the
recurrence plot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
8.1 The rocking spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
8.2 Elements of a typical collapse orbit; angle, velocity and acceleration. . . . . . . 200
8.3 Comparison of phase orbits for collapsed and no collapsed orbits . . . . . . . . 201
8.4 Mechanical energy of a typical collapse orbit . . . . . . . . . . . . . . . . . . 201
8.5 Dependence of the probability of collapse with slenderness ratio . . . . . . . . 203
8.6 Dependence of the probability of collapse with the frequency parameter . . . . 203
8.7 Dependence of the probability of collapse with the damping coecient . . . . . 204
8.8 Probability of collapse as function of . . . . . . . . . . . . . . . . . . . . . . 204
xix
8.9 Comparison of numerical and experimental probability of collapse . . . . . . . 205
8.10 Bounded states in the eective potential . . . . . . . . . . . . . . . . . . . . . 208
8.11 Modied Poisson model for the probability of collapse . . . . . . . . . . . . . 212
8.12 Fokker-Planck model for the probability of collapse . . . . . . . . . . . . . . . 214
B.1 Denition of rotations about an axis . . . . . . . . . . . . . . . . . . . . . . . 226
C.1 Contour of integration for Eq. C.9 . . . . . . . . . . . . . . . . . . . . . . . . 230
C.2 Numerical test for the phase in Eq. C.17 for = 10
5
and = 10
2
. . . . . . . 232
E.1 Reference system for data acquisition . . . . . . . . . . . . . . . . . . . . . . 238
E.2 System data acquaintance. LED-Camera-Mirror system (left) and general view
(right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
E.3 Details of Data acquisition devices. LEDS . . . . . . . . . . . . . . . . . . . . 239
E.4 Period denition for parameter-tting . . . . . . . . . . . . . . . . . . . . . . 240
F.1 Noise induced by the shaking table . . . . . . . . . . . . . . . . . . . . . . . . 243
F.2 Induced noise: harmonic motion tests . . . . . . . . . . . . . . . . . . . . . . 243
F.3 Induced noise: random motion tests . . . . . . . . . . . . . . . . . . . . . . . 244
F.4 Fourier spectrum of table transfer function . . . . . . . . . . . . . . . . . . . . 244
F.5 Specimen 1. Horizontal displacement and torsion angle for hanning sine of
3.3Hz and 10mm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
F.6 Specimen 2. Horizontal displacements and torsion angle for constant sine of
4Hz and 8mm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
F.7 Specimen 2, maximum torsion angle for random motion . . . . . . . . . . . . 247
F.8 Specimen 4, Horizontal displacement and torsion angle for run-down sine test
of 7.0 0.7 Hz and 2 mm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
F.9 Specimen 3. Horizontal displacement and torsion angle for constant sine of 5Hz
and 2mm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
F.10 Specimen 3. Torsional vibration . . . . . . . . . . . . . . . . . . . . . . . . . 249
xx
List of Tables
4.1 Hinge vectors for a system of N stacked blocks . . . . . . . . . . . . . . . . . 52
6.1 Types and number of tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.2 Mechanical properties of granite stone . . . . . . . . . . . . . . . . . . . . . . 74
6.3 Stone dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.1 Specimen 2, absolute value of maximum rocking angle (
o
) for harmonic motion. 100
7.2 Experimental linear correlation coecient for m
n
and m
n+1
. Free rocking . . . . 113
7.3 Experimental linear correlation coecient for maxima under harmonic loading.
Eect of external amplitude . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.4 Experimental linear correlation coecient for maxima under harmonic loading.
Eect of external frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.5 Experimental linear correlation coecient for maxima under random loading.
Specimen 3 under dierent PGA values . . . . . . . . . . . . . . . . . . . . . 118
7.6 RQA estimators for harmonic forcing. Periodic modes . . . . . . . . . . . . . 183
7.7 RQA estimators for harmonic forcing. Undamped periodic modes . . . . . . . 183
7.8 RQA estimators for harmonic forcing. Quasi-periodic modes . . . . . . . . . . 184
7.9 Comparison of the RP estimators for the three main types of bounded orbits in
the harmonic forcing regime . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
8.1 Number of observed overturns for specimen 3 . . . . . . . . . . . . . . . . . . 199
E.1 Dynamical parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
xxi
F.1 Specimen 2, maximum torsion angle (
o
) for harmonic motion. . . . . . . . . . 247
F.2 Specimen 4, maximum torsion angle (
o
) for harmonic motion. . . . . . . . . . 248
F.3 Geometrical characteristics of the specimens. . . . . . . . . . . . . . . . . . . 250
xxii
Part I
Introduction and Critical Review of
Classical Theory
1
Chapter 1
Introduction
1.1 Rocking Motion and Historical Constructions
Earthquake hazard in the Iberian Peninsula
This work started to be written in the celebration of the 250 anniversary of the historical 1755
Lisbon earthquake; one of the most devastating episodes in the history of natural disasters.
That earthquake aected not only Portugal but also the majority of the Iberian Peninsula and
Europe [1]. The magnitude of loss of lives was probably increased due to a fatal coincidence;
the misfortuned day of the shock -November 1
st
- most of the people were at temples celebrating
the All Saints Day, a traditionally religious event in Spain and Portugal.
The XVIII century elite of intellectuals was so impressed, that a number of contributions to
Philosophy, Art, and Science found their birth at that time related to the Lisbons catastrophe.
It is known that even the great philosopher Kant himself spent a long time studying all the sur-
veys he could obtain in an attempt to develop a model for earthquake risk. In fact, Lisbon 1755
earthquake can be considered as the birth of modern geophysics since, for the rst time, many
thinkers and scientists started to consider the problem of earthquake hazard as an important
problem.
2
1.1. Rocking Motion and Historical Constructions 3
Figure 1.1: Ruins of Convento do Carmo at the city of Lisbon, Portugal.
The other setback of the 1755 earthquake was its eect on monuments. Hundreds of reports
radiating from cities all over the peninsula appeared describing more or less damage in their
principal buildings. Ranging from total collapse, as it was the case of Convento do Carmo in
Lisbon (Fig. 1.1), to partial damage at the bells tower of the Cathedral in Salamanca (Spain),
hundreds of buildings were aected by the ground shake [1].
Spain and Portugal are considered as lowto moderately-high seismicity regions (see Fig. 1.2 and
[2]), with special recommendations found for non-engineered construction such as old masonry
buildings. For these cases specic provisions would be needed in order to take into considera-
tion the particular dynamical performance of masonry buildings.
For instance, the Spanish code (NCSE-02) [3] does not allow dry-joint masonry structures if the
design peak ground acceleration is greater than 0.04g for those buildings with an importance
considered as moderate and signicant (see [3]).
1.1. Rocking Motion and Historical Constructions 4
Figure 1.2: Iberian Peninsula hazard map. Mean value of horizontal ground acceleration (m/s
3
)
for a 975 years return period. Probability of exceedance of 5% within 50 years. [2]
It is worth to notice the lack of criteria for the quantication of risk in existing old masonry
structures, such as monuments, despite the recent European eorts [4], and US eorts [5].
On the other hand, conservation and repair of historical buildings has emerged as a recent and
active eld of interest for the scientic and technical communities with signicant importance
for the society in general. As claimed in the Venice Charter [6]:
"People are becoming more and more conscious of the unity of human values
and regard ancient monuments as a common heritage.
The earthquake hazard, to which architectural heritage is exposed, has proved to be extended
and generalized all over the world, being at the same time one of its most devastating agents.
Nevertheless, the assessment of the integrity and conservation of historical structures is far from
straightforward. Safety level quantication and the achievement of a full understanding of ma-
sonry structures behavior is still a challenge.
Portugal and Spain are very rich countries in architectural heritage. Within the set of penin-
sular structures under protection, special regard deserve Roman and Medieval architecture,
1.1. Rocking Motion and Historical Constructions 5
Arabesque palaces and Mosques, Gothic Cathedrals, Renascence Churches and Baroque Churches
and palaces.
Despite the increasing demand of new methods for damage quantication and safety assess-
ment, the so-called Rocking Motion dynamics (RM), which is dominant for masonry structures,
has been shown to be extremely dicult to handle.
The rocking motion problem
A rigid body placed on a foundation is said to be in RM if its center of rotation changes in-
stantaneously between dierent contact points when passing through its equilibrium position.
There is an energy loss associated to each impact with the foundations at every cross through
the equilibrium point.
One common mistake in earthquake engineering stems from the incorrect assumption that RM
only appears for single block structures or under strong near-source ground motions. This
wrong perception of the problem is not only erroneous but also risky, as it will be shown along
this work.
The key point to be highlighted is that it is not necessary for the whole building to undergo
rocking -which could be also the case- for RM to be present. In fact, masonry buildings fail in
a system of macro-blocks, with discrete cracks, being RM present at every interface if sliding
does not occur because, the system lacks any other mechanism to dissipate energy. In this
case, the relevant energy is not elastic but gravitational and erroneous results can be obtained
when the frequencies of the structure are derived [7], [8].
As a consequence, even if there is micro-rocking at joints (cracks), the overall damping mecha-
nism and, hence, the overall dynamics, will be dierent.
Nevertheless, only a few new approaches take into account the presence of RM when dealing
with masonry structures (see section 1.2).
1.1. Rocking Motion and Historical Constructions 6
The masonry material
A point of controversy for many researchers is which model should be adopted for a masonry
building when subjected to dynamic loading [9]. The key of this polemic is the particular
mechanical behavior of masonry material.
Masonry material is a heterogeneous, very complex and, usually, anisothropic material [10]. Its
behavior depends on the nature of units, stones, bricks, etc, and the kind of joints, usually lled
with lime mortar, clay, or others. Its strength depends on the dimension of single elements, the
proportions of constituents, the arrangement in the structure, the direction and of joints etc. (see
Fig. 1.3). However, the mechanical behavior of the dierent types of masonry exhibits generally
Figure 1.3: Dierent kinds of stone masonry: (a) rubble masonry; (b) ashlar masonry; (c)
coursed ashlar masonry.
the common feature of a very low tensile strength. This property is so important that the form
and dimensions of ancient constructions was determined according to it [11].
On the other hand, despite of the apparent simplicity of the RM dynamics, a big wealth in
behavior emerges even for the simplest single-block structures.
Amongst other topics; quasi-periodic response, chaotic response and heteroclinic bifurcations
have been found for the single block system, [12]-[15]. These particular features of the RM
dynamics have highlighted the necessity of a careful decision about which dynamic framework
should be used for a correct modeling of a masonry building under earthquake action.
1.2. Models for Masonry Dynamics 7
1.2 Models for Masonry Dynamics
When an assemblage of blocks undergoes horizontal actions, the system develops a mechanism
in order to react to that loading. If friction is enough to prevent sliding, the mechanism will
consist on the formation of fracture lines acting as instantaneous axes of rotation.
Although sliding mechanisms might take place in masonry construction, namely at higher lev-
els of base acceleration [16], collapses also occur due to the formation of hinged mechanisms.
The division of the whole body under analysis into rigid macro-blocks, when it is exposed to
severe loading, such as earthquakes, is a natural consequence of the hypotheses of zero tensile
strength. Therefore, models that consider large rigid assemblies, connected each other by ro-
tation ones and rocking at these axes can reproduce the behavior of a large group of masonry
constructions under earthquakes.
At the present time, the reference method for determining the dynamic performance of a ma-
sonry building is the modal response analysis, using a linear-elastic model of the structure,
despite the recent development in push-over nonlinear analysis. However, as stressed above,
consideration of linear elastic behavior for historical masonry construction is, generally, not
recommended [17].
A powerful alternative to this basic method is the nonlinear time history analysis. The obvious
diculty in using direct integration of the dierential equations of motion, using base accelero-
grams, is that the time requirements render most analysis non-feasible. It is noticed that a
minimum of ve accelerograms should be used (only applies in few countries). This results in
twenty dierent analysis, without considering the horizontal components of the seismic action
acting simultaneously and the vertical component of the seismic action.
As suggested by Part 3 of Eurocode 8 [18], it might be more appropriate that the analysis be
carried out using approximate static nonlinear methods.
Static methods
In the static methods, or the so-called push-over analysis the seismic eect is dened by the
base shear coecient, termed as seismic coecient, which denes the percentage of the total
1.2. Models for Masonry Dynamics 8
weight of the building that must be considered as a horizontal force, also total, applied to the
structure.
For modern buildings, existing codes adopt detailed procedures to calculate the seismic coef-
cient applicable for each case, and as a function of the seismic zone, soil type, fundamental
frequency of vibration for the structure, and the available ductility and damping. These pro-
cedures have been extensively calibrated against results obtained by the general methods, but
historical buildings possess characteristics considerably dierent from the ones used to cali-
brate the codes. For areas of signicant seismic hazard, the seismic coecient can be assumed
to vary between 0.1 and 0.3 [19].
The application of this approach is still time consuming and requires advanced engineering, able
to perform sophisticated nonlinear analysis. International provisions as Eurocode 8 or FEMA
also introduce the behavior factors in order to hold the plastic behavior out of the elastic range.
For instance, Eurocode gives a value of 1.5-2.5 for ordinary unreinforced masonry and 2.0-3.0
for conned masonry [4]. All these quantities must be revised and explained if we are to provide
a better understanding of the masonry structures response during earthquakes.
Limit analysis
Limit analysis combines, on one hand, sucient insight into collapse mechanisms, ultimate
stress distributions (at least on critical sections) and load capacities, and on the other, simplicity
to be cast in a practical computational tool. Another appealing feature of limit analysis is the
reduced number of necessary material parameters, given the diculties in obtaining reliable
data for historical masonry.
1.2. Models for Masonry Dynamics 9
The applicability of limit analysis theory to un-reinforced masonry structures modeled as as-
semblages of rigid blocks interacting through joints rests on some basic hypothesis:
limit load occurs at small overall displacements,
no tensile strength,
perfectly plastic shear failure at joints,
hinging failure mode at joints occurs with compressive force being independent of the
rotation.
The rst hypotheses is reasonable for most cases. On the other hand, the low tensile strength and
the quasi-brittle tensile failure of masonry justify the second assumption. The third hypothesis
is conrmed by experimental results. The last hypothesis could be questioned when crushing
occurs, but the agreement of numerical results with experimental ones, justify this hypothe-
sis. Livesley [20] was the rst author to apply the limit analysis theory to rigid block arches,
by means of the static theorem. More recently [21] applied the kinematics theorem to the same
problem, aiming at analyzing masonry arch bridges. These authors considered that masonry has
an innite compressive strength, and that the shearing failure, governed by the Coulombs law,
is associated. Begg and Fishwick [22] have released the restriction of associated ow rule or
normality condition for masonry arches, including sliding shear failure. Baggio and Trovalusci
[23] addressed also this aspect in a general formulation for rigid block limit analysis. Ordua
and Loureno have recently improved the solution procedures for 3D problems.[24],[25]. It is
noted that extending the formulation to non-associated ow, results in a nonlinear mathemat-
ical problem of a signicantly larger size, in comparison with the lower size linear problem
resulting from the classical theory. Furthermore, that nonlinear problem is very hard to solve.
An important contribution from the mathematical programming viewpoint is that of Ferris and
Tin-Loi [26].
Finite element methods
Nonlinear nite element approaches are a possible choice for masonry structures assessment.
Both advanced continuous, anisotropic based models, and discrete (micro-) models for masonry
1.2. Models for Masonry Dynamics 10
structures have been developed in the last decades, e.g. [10],[27].
Nevertheless, the drawback of using nonlinear nite element analysis in practice might include:
requirement of adequate knowledge of sophisticated nonlinear processes and advanced
solution techniques by the practitioner,
comprehensive mechanical characterization of the materials,
large time requirements for modeling, for performing the analyses themselves, with a
signicant number of load combinations, and for reaching proper understanding of the
results signicance.
Of course, for special cases, as complex, important or large structures, nonlinear analysis should
not be disregarded as an analysis tool.
Linear-elastic analysis can be assumed a more practical tool, even if the time requirements of
modeling are similar. Such an analysis fails to give an idea of the structural behavior beyond
the beginning of cracking. Due to the low tensile strength of masonry, linear elastic analysis
seem to be unable to represent adequately the behavior of historical constructions.
Discrete element methods
Discrete-element based algorithms (DEM) are a further step towards the correct modeling of
rigid blocks with joint interfaces [28], [29]. The typical characteristics of discrete element
methods are:
a) the consideration of rigid or deformable blocks (in combination with FEM),
b) connection between vertices and sides / faces,
c) interpenetration is usually possible,
d) integration of the equations of motion for the blocks, (explicit solution) using the real
damping coecient (dynamic solution) or articially large (static solution).
The main advantages are an adequate formulation for large displacements, including contact
update, and an independent mesh for each block, in case of deformable blocks.
The main disadvantages are the need of a large number of contact points required for accurate
1.2. Models for Masonry Dynamics 11
representation of interface stresses and a rather time consuming analysis, especially for 3D
problems. In order to overcome this drawbacks, a new research is being developed in an attempt
to use the Rocking formulation and DEM models together [30].
Figure 1.4: The curved-base block

R
2b
2h
o
o

c.m.
Figure 1.5: RM model
Figure 1.6: Modal analysis method Figure 1.7: DEM method
Motivation for RM analysis. The problem of frequencies
Figures 1.4-1.7 depict four models for a single rigid block dynamics. Figure 1.6 and Fig. 1.7
show standard modal analysis method and DEM approaches, respectively. Figure 1.5 shows the
RM model. Finally, Fig. 1.4 represents an interesting case; a block with a certain curvature at
its base.
1.3. An Overview of the Existing Works related to RM 12
As it is well known, from continuum based models it is possible to obtain periods which are
independent of the amplitude of the oscillation for small displacement regime. However, this
is not the case for the RM model, where periods strongly depend on the amplitude (external
energy). This fact is due to a singularity present in the RM potential which prevents series
expansion about the equilibrium point. As it will be shown along this work, this fact turns
RM dynamics into a completely dierent problem than that derived from standard continuous
models.
The case of the curved block can be consider as a transition between discrete and continuous
models although for this case, as in RM, the relevant energy is still gravitational and not derived
from body deformation. The periods for this block are independent of the amplitude for small
oscillations (see appendix A for details). Nevertheless, as the curvature of the base tends to zero
the period starts to diverge.
1.3 An Overview of the Existing Works related to RM
The study of the rocking motion (RM) of rigid objects has been a eld of increasing interest
to researchers for over a century, specially in connection with standing architectural objects
(columns, towers, storage tanks, etc.) subjected to ground motion earthquakes, machine vibra-
tions, etc. or surveillance and operational equipment on board of oshore structures.
Early modern studies of RM date back to the works by Milne [31] in 1881, where the over-
turning of tombstones and monumental columns in Japan was used as an estimator of the peak
acceleration of an earthquake excitation.
However, the classical theory for a planar rocking block resting on a rigid foundation was rst
brought by Housner [7] in an eort to explain a surprising empirical evidence. During the
1960 Chilean earthquakes, several golf-ball-on-tee water tanks remained standing while more
seemingly stable structures in the immediate vicinity collapsed by overturning. In his funda-
mental paper, Housner reported the most important characteristics of rocking phenomenon of
the so-called Inverted Pendulum Structures.
1.3. An Overview of the Existing Works related to RM 13
More recently, dierent authors analyzed the behavior of a block subjected to dierent earth-
quake inputs while others focused their works in the stability of RM dynamics under harmonic
excitation. In this latter context, it has been shown how RM can become chaotic for some values
of the parameters involved.
An overview of the existing related works leads to a classication based on four main criteria:
type of structure under study,
inclusion of experimental research,
type of analysis performed,
consideration of additional eects, such as bouncing or sliding.
The rst division concerns the type of structure under study; single degree of freedom or mul-
tiple degree of freedom systems. Although a number of relevant contributions in the eld of
multiple-body systems have appeared in the last decades, there is still a gap concerning rocking
mechanisms.
Lipscombe [9], Augusti [17], Winkler [32], Psycharis [33] and Sinopoli [34] tackled the prob-
lem of an ensemble of rigid blocks. The structures studied were columns, un-reinforced ma-
sonry walls, and simple closed-chain kinematic mechanisms.
Lipscombe [9] focused on the dynamics of two blocks through a Lagrangian formalism. How-
ever, that Lagrangian was still splitted in two sets of equations rendering equations which were
not easy to handle. The works of Doherty et al. [35] and Felice [36] dealt with dierent col-
lapse conditions for un-reinforced masonry walls. From another perspective, RM can also be
considered as a non-holonomic constraint problem. In this eld, the works by Marsden [37] and
Leon [38] are closely related to the RM problem. However, in their present form, those works
may nd little application in earthquake engineering.
At the present time, there is still a lack of proper experimental RM research. Aslam et al. [39]
reported the empirical basis for a far-reaching discussion about which parameter values should
be assumed in the model. They also found diculty in reproducing the response under the same
conditions. This fact, which was consequence of the high nonlinear behavior of the system, led
to a cascade of contributions reporting the chaotic nature of the problem.
1.4. The Present Contribution. Objectives and Scope 14
Wong et al. [40] conducted an exploration of the main types of orbits, reporting new types not
previously found.
Regarding the type of analysis performed, two main groups are found in literature. On one hand,
the approach is more probabilistic oriented, with a clear focus on earthquake engineering and
applications. On the other hand, the main eort is conducted to the stability of orbits over the
parameter space. The works by Spanos [14], Yim et al. [41], Makris et al. [42], Liberatore [43]
and Ishiyama [44] belong to this classication. These contributions have a more applied prole
and they provide criteria for the probability of block overturn under parameter variations.
As stressed above, nonlinear dynamics [45] is an inherent characteristic of RM dynamics.
Works dealing with nonlinear dynamics consider harmonic pulses as external loading since,
as it will be shown, analytical expressions for the solution can be derived in this case. Theoret-
ical predictions in the parameter space and nonlinear techniques such as Melnikov have been
used in this context.
In this regard, Hogan [12, 13], [46], Yim et al, [15, 47], and Sinopoli [34] determined under
what conditions the system could become unstable and, eventually, chaotic.
Finally, dierent eects such as sliding, bouncing or combinations of these both RM have been
also considered. Augusti [17], Shenton [48] and Jeong [49] include sliding as part of the prob-
lem. They provide conditions for the initiation of sliding under certain conditions. Coupling
between these mechanisms and RM is also examined.
1.4 The Present Contribution. Objectives and Scope
The main objective of this work is to provide an adequate basis for the analysis of rigid block-
assemblies dynamics with special attention to old masonry structures. The new formulation
proposed can complement standard approaches reported in section 1.2, specially, limit analysis
and DEM. A novel formulation for the rocking motion of rigid block assemblies is introduced
by means of generalization of the results found [50] and [51]. For that analysis, the basic
conguration space is transformed into a complex Riemannian manifold with an Hermitian
inner product.
1.4. The Present Contribution. Objectives and Scope 15
Concerning applications, the probability thresholds of collapse for slender structures are found.
Conditions for the stability of the four-hinged mechanism of a masonry arch are also examined.
In order to validate the analytical development, experimental research program was carried out
at the National Laboratory of Civil Engineering (LNEC), (Lisbon, Portugal). This research is
concerned only with no-sliding mechanisms. Although the proposed formalismfor many blocks
is not limited to two dimensions, only applications to planar congurations are included.
The organization of this work is as follows. Part I presents the RM problem. The so-called
"classical theory" is discussed and critically reviewed. Part II proposes a formulation for many
degree of freedom systems. Part III deals with the dynamics and stability of single degree of
freedom systems; both theoretical results and extensive numerical study are reported. Finally,
Part IV presents some practical applications to earthquake engineering and structural dynamics.
Appendices complement the work in those details which, although considered fundamental, can
be omitted along the text in the seek of a clearer exposition of the ideas.
Due to the extensive use of mathematical expressions and results along the entire text, it was
considered appropriate to include the most important derivations and formulas in appendix B.
Chapter 2
Critical Review of the Classical Theory
The Piecewise theory (PWT), also called classical theory, is revisited in this chapter for com-
parison and reference.
2.1 PWT Equations of Motion
The main hypothesis assumed in the PWT are that:
1. the block and its base are rigid,
2. there is no sliding,
3. in-plane (2D) motion is only considered (out-of-plane (3D) behavior is neglected),
4. the impact of the block, during the rocking, is not elastic. Thus, the block cannot jump.
This means that exists, always, at least one contact point between the RB and its base.
5. Angular moment conservation, before and after impact, holds.
Let M and I be the mass and moment of inertia
1
of a rigid rectangular block of height and
width 2h and 2b, respectively (see Fig. 2.1. If the base is subjected to an acceleration a(), the
block will no longer be in an inertial frame of reference and it will experiment inertial forces
distributed all over its mass. For an observer traveling with the foundation, the eect is a force
concentrated at the center of mass of magnitude Ma().
1
Measured from O or O

(see Fig. 2.1), I is simply I


4
3
MR
2
.
16
2.1. PWT Equations of Motion 17

R
2b
2h
o
o

c.m.
Figure 2.1: Geometry and notation for the rocking block.
The angle that the block forms with the foundation is denoted by and the time is dened
through the symbol . Positive angles are dened counterclockwise. Time-dierentiated quan-
tities are indicated by an apostrophe

d/d.
From DAlemberts principle, and by taking moments about O and O

, the following equation


is obtained.
I

MRa() cos( ) MRg sin( ) = 0 (2.1)


where the sign refers to the domains > 0 and < 0, respectively and g is the acceleration of
gravity. Other parameters, like R and , are dened according to Fig. 2.1.
Equation 2.1 can be set in a more compact form by using the absolute value and the signum
function
I
d
2
d
2
|| + MgRsin( ||) sign()Ma()Rcos( ||) = 0. (2.2)
The quantity || is bounded between 0 and /2 since, clearly || > /2 has no physical meaning.
Therefore, natural barriers appear for :
0 ||

2
. (2.3)
In addition, the angle is also bounded between the same limits. The case = 0 corresponds
to an innite thin block whereas =

2
represents a block with innite width.
2.1. PWT Equations of Motion 18
For harmonic forced excitations, the following form for a() is commonly assumed,
a() = g cos( + ). (2.4)
where , and are the non-dimensional amplitude, circular frequency and phase of the
external action, respectively.
In the following, a non-dimensional angle x, time t, and frequency are adopted using and
parameter p as p
_
MgR/I,
x

(2.5)
t p

p
.
Here it is noted that, for a rectangular block p reads p =
_
3g/4R.
If dierentiation with respect to t is referenced as in Newtons notation: x dx/dt, Eq. 2.2,
with the denition for a() of Eq. 2.4 and the normalization given in Eqs. 2.5 becomes
| x| + sin((1 |x|)) sign(x) cos((1 |x|)) cos(t + ) = 0. (2.6)
Piecewise linearized theory
For slender blocks and small angles it is natural to assume

cos((1 |x|)) 1,
sin((1 |x|)) (1 |x|).
From Eq. 2.6, it follows
| x| |x| + 1 sign(x) cos(t + ) = 0. (2.7)
Assuming the phase of the base acceleration being equal to zero and multiplying the preceding
equation by sign(x) one obtains:
x x + sign(x) = cos(t). (2.8)
2.2. PWT Model for Impact 19
This equation is well-known in the related literature. For free rocking regime ( = 0) it coincides
with the Housner equations for slender blocks [7], given by
x x 1 = 0. (2.9)
If t

represents the time of impact, the solutions for each sign of x for Eq. 2.8 are

x
(+)
(t) = A
(+)
cosh(t t

) + B
(+)
sinh(t t

) cos(t) + 1,
x
()
(t) = A
()
cosh(t t

) + B
()
sinh(t t

) cos(t) 1.
The superscript (+) is valid for x > 0, and the superscript (-) for x < 0. The parameter is
dened by = /(1 +
2
) while A
()
and B
()
are constants of integration depending on the
initial conditions. For initial conditions {x(t

) = x
0
, x(t

) = v
0
}, the values for A
()
and B
()
are
A
()
= x
0
1 + cos(t

), (2.10)
B
()
= v
0
sin(t

).
For free rocking ( = 0) and with initial conditions,
{x(0) = x
0
> 0, x(0) = 0}, (2.11)
Housner computed the period by setting x = 0 in the rst of Eqs. 2.9 and then solving for
the time, which is one fourth of the period of the orbit. Thus, if T
H
denotes the period of
the classical piecewise theory (the superscript "H" here is in reference to Housner), its nal
expression results in is
T
H
= 4 cosh
1
_
1
1 x
0
_
. (2.12)
2.2 PWT Model for Impact
As stressed above, every time the block passes through the position x = 0 there is a reduction
of energy and, thus, sudden changes of velocity at the point x = 0 take place. The traditional
theory assumes that if x
b
and x
a
are the velocities just before and after each impact, they can be
related through:
x
a
x
b
= (2.13)
where, is a constant quantity dened as the coecient of restitution.
2.2. PWT Model for Impact 20
Although the inclusion of this coecient points to the existence of an impulsive interaction, no
mention is found in previous works regarding the form and magnitude of such force.
Housner [7] provided a theoretical value of under the assumption of angular momentum con-
servation about O or O

when the impact takes place. By equating the angular momentum just
before and after the impact, equation
I
b
2bMRsin()
b
= I
a
, (2.14)
holds, which, after simple manipulation, renders a formula depending only on the geometry of
the block
= 1
3
2
sin
2
. (2.15)
Several discrepancies have been found between the theoretical value given in Eq. 2.15 and the
experimental determinations of [39, 43]. Due to these deviations, it has been suggested to
assume either that the impact is not perfectly inelastic or to adopt empirical measured values
for [39, 40]. In any case, the coecient of restitution procedure seems to cast several doubts
about its validity.
The classical formulation for the one block free rocking motion in its linearized version is
constructed by adding Eq. 2.9 to Eq. 2.13

x x + 1 = 0 x > 0
x x 1 = 0 x < 0,
x(t
a
) = x(t
b
)
where t
a
and t
b
represent the times before and after the impacts, respectively.
Because for free RM, no dissipative forces other than impulses act on the block, energy is con-
served between two consecutive impacts. Since the change in the kinetic energy is proportional
to the square of , the relation between the two energies, before and after, the impact is
E
a
E
b
=
2
. (2.16)
This formula is directly generalized for an arbitrary number of impacts.
From an initial level, jumps to lower levels occur at each impact (see Fig. 2.2 a)) and the height
2.3. Summary 21
a) Non-dimensional energy as a function
of time for: e
0
= 0.375, =0.8 and initial
conditions: x
0
= 0.5, v
0
= 0.0
b) Energetic levels obtained from Eq. 2.17
for: =0.5 and 0.8 with e
0
= 0.375
Figure 2.2: Energetic map for a RM block
of each level is given by the expression:
E
N
= E
0

2N
, (2.17)
being E
0
the initial energy (function of the initial conditions) and N the impact counter, which
also labels the energetic levels. Fig. 2.2 b) shows two typical energy diagrams for = 0.5 and
= 0.8 obtained from Eq. 2.17 with e
0
= 0.375.
2.3 Summary
In this chapter the main ideas of the so-called classical theory of rocking motion have been
exposed. The most relevant features of this formulation are the piecewise nature of the equations
and the impulsive force acting at every impact.
In addition, the impact, in its present form, is introduced as an ad hoc mechanism, independent
of the equations of motion.
These facts clearly make it dicult to generalize the formalism. As it will be shown in the next
section, an alternative approach is possible if the angle is allowed to be a complex quantity.
2.3. Summary 22
The success of the novel approach is that the equations of motion split into an equation for the
modulus and one for the phase. This new perspective allows not only a deeper insight into the
physics of the problem but also a generalization to a larger number of blocks.
Part II
Formulation for Multiple Degree of
Freedom Systems
23
Chapter 3
Geometrization of RM Multi-Block
Dynamics
In this chapter the formulation for an arbitrary number of rocking blocks is presented. Expres-
sions for the corresponding kinetic and potential energies are derived.
3.1 Outline of the Problem
The dynamics of rocking block assemblies can be understood as the coupling of two main
concepts, namely: collapse mechanisms of a given body and coupled rotations between bodies.
In a schematic way:
Dynamics = Mechanisms Rotations
In what follows, the following assumptions are made:
1. Bodies are solid undeformable polygons with a given number of faces and edges,
2. No sliding is allowed.
3. Each of the edges of the body constitutes a potential axis of rotation (rotations are there-
fore considered to be allowed only about these edges, when one of the axes develops into
a fracture line),
4. Fracture lines in 3D (axes) correspond to hinges in 2D, or planar congurations.
24
3.2. Set-up of the Conguration Space 25
In the model, it is assumed that the domain consist of the disjoint union of N sets, each one
corresponding to one of the components,
=
N
_
n=1

n
(3.1)
Typical problems in continuum mechanics such as intersections and penetrability are then ig-
nored.
The procedure to reach the correct expressions for the kinetic and potential energy of a given
problem consists of the following steps:
1. The kinetic and potential energies are rst found in the Euclidean space. The aim is to
cast the kinetic energy T into its standard form:
T =
1
2
a
ik
q
i
q
k
where the sum is extended over the degrees of freedom q
i
, q
i
are the generalized velocities
and a
ik
are, in general, functions of the q
i
. These expressions are still uncoupled with
mechanisms.
2. A complex manifold with inner Hermitian product is built as the basic conguration
space.
3. Then a metric tensor is proposed in the complex-manifold of the conguration space.
This metric carries the mechanism coupling.
4. Finally, the mechanism-coupled kinetic and potential energies are derived within this new
geometry.
3.2 Set-up of the Conguration Space
The starting space consist of N copies of the Euclidean q-dimensional space R
q
:
R
q
R
q
. . . R
q
= R
qN
(3.2)
where q is the dimension of the space under consideration (in this work q equals two or three).
3.2. Set-up of the Conguration Space 26
Reference frames
Although the nal expressions derived are reference-free, it is convenient to dene a global
reference frame by the pair (O, e
i
), i = 1..q, where O denotes a point of the associated Euclidean
ane space called "origin". This frame is simply referenced as O. Figures 3.1 and 3.2 illustrate
the model. For each n-subspace, the chosen system of reference consist of the pair (h
n
, u
n
),
Figure 3.1: Refence frames for rigid rota-
tions of two-block assemblies
Figure 3.2: Rotation about three fracture
lines.
where h
n
denotes the hinge of body n and the vectors u
n
form a mobile basis tied to the solid
1
.
This frame will be labeled as n-frame. The bases of each n-frame are sets of three unitary
orthogonal vectors built according to the right-hand rule. The third element of the basis, u
3
,
represents the rotation axis (fracture line).
With this notation for the unitary base vectors, superindexes n denote the body at which that
basis is tied to and sub-indexes label the vector of the basis.
As it is know [52], the position of every point of body n can be determined by knowing the
position of its hinge h
n
and the orientation of its non-inertial basis {u
n
(t)} with respect to O.
1
In this expression "tied" means moving with the body
3.2. Set-up of the Conguration Space 27
Conguration vectors
A conguration vector of this space is dened as the N-tuple:
r = (r
1
, r
2
, . . ., r
N
), (3.3)
where each r
n
represents the position of a particle of body n with respect O.
Basis vectors
As it is customary in multiple-body theory, basis vectors of R
qN
are built as the tensorial product
of the basis vectors of each subspace.
Material vectors
The material vector x is dened as:
x = (x
1
, x
2
, . . . x
N
), (3.4)
where the vectors x
n
, dened in the n-frame, connect the hinge of body n with any material
point of that body.
Hinge vectors
The connection through axis (hinges) of rotation is accomplished through the

d vector, dened
as:

d = (

d
01
,

d
12
, . . .

d
N1N
). (3.5)
Here, each vector

d
nn+1
, dened in the n-frame, connects a hinge (axis) of body n with a
hinge(axis) of body n + 1. The rst vector

d
01
connects the origin of the reference system
with the hinge of body 1.
It is always possible to choose a reference frame where

d
01
= 0, and hence:

d = (

0,

d
12
, . . .

d
N1N
). (3.6)
3.2. Set-up of the Conguration Space 28
Projection operators
The projection operator for the n-subspace is dened by:

P
n
=
N

m=1

n
m

I. (3.7)
Its matrix in R
qN
is given by:
P
n
=
N

m=1

n
m
I
qxq
, (3.8)
where I
qxq
is the square identity matrix of dimension q.
Density operators
The density operator is dened by:
=
N

n=1

P
n
, (3.9)
where
n
is the density of body n.
Rotation operators
The rotation operator is dened through the expression:

R =
N

n=1
e

n
. (3.10)
Here the operators are the so-called generators of the rotations (see appendix B). The matrix
of

R is given by:
R =
N

n=1
e

n
(3.11)
Chain Operators
The following

C operators are introduced by means of their matrices C
mn
in R
2q
, given by:
C
mn
=

I
q
x
q
, i f m > n
O
q
x
q
, otherwise.
(3.12)
These matrices are strictly upper triangular of dimension (qN)
2
and, hence, it can be shown that
these operators are nilpotent.
3.2. Set-up of the Conguration Space 29
Expression of the conguration vectors under rotations
In order to derive the kinetic and potential energy expressions for arbitrary rotations it is rst
necessary to reach the expression of the conguration vectors under a general transformation.
The conguration vector r for an arbitrary rotation is dened as:
r =

R(x) +

C

R(

d) (3.13)
Notice the interesting coupling; whereas the material vectors are simply rotated, the hinge vec-
tors are also coupled through the

C operator. It should be stressed that in general

C and

R do
not commute.
In matrix form:
r = xR + dRC (3.14)
The preceding equations are, in fact, the denition of the conguration space. Eq. 3.13 denes
a point transformation through:
r = r(
1
, . . . ,
N
) (3.15)
This implies a change of R
2q
to S
N
. The topology of the new space is therefore equivalent to a
N-torus.
Kinetic energy in S
N
The kinetic energy of the system is dened by
T =
1
2
_

d <

r, (

r) > (3.16)
where d is a certain measure over the domain . By labeling /
k
=
k
it holds:

r =

k
r (3.17)
<

r, (

r) >=
N

i, j=1

j
<
i
r, (
j
r) > (3.18)
3.2. Set-up of the Conguration Space 30
By using the expression of the density operator:
<
i
r, (
j
r) >=
N

n=1

n
(
ir
)P
n
(
j
)
T
(3.19)
and then,
T =
1
2
N

i, j=1

j
N

n=1
_

d
n
(
i
r)P
n
(
j
r)
T
. (3.20)
The generalized moments of inertia are now dened as:
J
i j

N

n=1
J
n
i j
, (3.21)
where the partial moments of inertia I
n
i j
corresponding to the n-subspace have been introduced:
J
n
i j

_

d
n
(
i
r)P
n
(
j
r)
T
(3.22)
By dierentiation of Eq. 3.14,
r
i
= X(
i
R) + d(
i
R)C (3.23)
Taking into account that the rotational operator for each k-subspace can be expressed as:
R

k
= e

k
, (3.24)
it is found that

i
R =
N

n=1

in
R

n
(3.25)
Now, by introducing the matrix
i
:

i
=
N

n=1

in
(3.26)
Eq. 3.25 can be reformulated as:
(
i
R) =
i
R (3.27)
Therefore (
i
r)P
n
(
j
t)
T
expands in three terms, which after manipulation render:
(
i
r)P
n
(
j
t)
T
=
in

jn
x
n
x
T
n
+ dA
n
i j
x
T
+ dB
n
i j
d
T
(3.28)
3.2. Set-up of the Conguration Space 31
where "x" is the matrix associated with the material vector x and
A
n
i j
= (
i
R)CP
n
(
j
R)
T
+ (
j
R)CP
n
(
i
R)
T
(3.29)
B
n
i j
= (
i
R)CP
n
C
T
(
j
R) (3.30)
Notice that A
n
i j
= A
n
ji
and (B
n
i j
)
T
= B
n
ji
. The quantities A
n
i j
and B
n
i j
can be recast into equations
which depend only on R, and not on its derivates
i
R. Algebraic manipulation leads to:
A
n
i j
=
i
RCP
n
R
T

T
j
+
j
RCP
n
R
T

T
i
(3.31)
B
n
i j
=
i
RCP
n
C
T
R
T

T
j
. (3.32)
It is straightforward to demonstrate that:
[R,
i
] = 0, i. (3.33)
The moments of inertia for each n-subspace are then given by:
J
n
i j
=
in

jn
_

d
n
x
n
x
T
n
+ dA
n
i j
_

d
n
x
T
+ dB
n
i j
d
T
_

d
n
(3.34)
Now assuming a uniform model for the density
n
= , n = 1..N, and performing the integra-
tions it is possible to obtain:
J
n
i j
=
in

jn
I
n
+ m
n
[dA
n
i j
x
T
c
+ dB
n
i j
d
T
], (3.35)
where:
m
n
is the mass of body n.
I
n
is the moment of inertia of body n in the n-frame.
x
c
is the matrix of the vector of the center of mass of body n in th n-frame.
Using the denitions:
A
i j

n
m
n
A
n
i j
(3.36)
B
i j

n
m
n
B
n
i j
, (3.37)
3.2. Set-up of the Conguration Space 32
the generalized moments of inertia can be cast into the expression:
J
i j
=
N

n=1

in

jn
I
n
+ dA
i j
x
T
c
+ dB
i j
d
T
, (3.38)
which can be further simplied by the aid of some denitions. First, the mass matrix is dened
as:
M

m
1
0 0
0 m
2
0
.
.
.
.
.
.
.
.
.
0
0 0 m
N

(3.39)
Second, three block matrices
2
are introduced: the inertia matrix I, the "A" matrix, and the "B"
matrix. The block-components of these matrices are the matrices I
i j
, A
i j
and B
i j
respectively.
With this material, the block matrix associated to J renders:
J = I + dAx
T
c
+ dBd
T
(3.40)
Finally, the vectors

S
N
are introduced:

(
1
, . . . ,
N
) (3.41)
and the standard bilinear form in velocities for the kinetic energy is found:
T =
1
2

J
T
(3.42)
Potential energy in S
N
If, in the O reference system, the unitary vector along the vertical direction is referenced as e
2
,
its tensorial extension to R
2N
is given by:

2
= e
1
2
e
2
2
. . . e
N
2
(3.43)
Then, the potential energy of the conguration can be expressed as:
U = g
_

d <
2
, (r) > (3.44)
2
A "block-matrix" is a matrix whose components a
i j
are also matrices.
3.3. Implementation of Mechanisms 33
U can be recast in matrix form with the aid of the "underlined" vectors (see appendix B):

P
n
(x) = x
n
(3.45)
and the following matrices:
Z
c

1
tr(M)
N

n
_

n
d
n
x
n
(3.46)
Q
1
tr(M)
N

n
m
n
P
n
(3.47)
where x
n
denotes the matrix associated to x
n
.
From Eqs. 3.9 and 3.13 and noticing that:
[

R,

P
n
] = 0, n (3.48)
through the denition of the F matrix:
F dRCQ + Z
c
R (3.49)
the potential energy results in:
U = g tr(M)F
2
(3.50)
where
2
is the matrix associated to
2
.
3.3 Implementation of Mechanisms
So far, no special consideration has been given in this analysis to the mechanisms
3
present in the
dynamics. If the equations of motion were to be derived from a Lagrangian; L=T-U, piecewise
expressions would be obtained, since the expressions for T and U are dierent depending on
which axis of rotation is materialized. These functions are still mechanism-dependent. As a
consequence, the elements of those functions, namely, moments of inertia, hinge and material
vectors will depend on the mechanism. The indexes of each mechanisms corresponding to
3
By "mechanism" it is understood the admissible kinematical collapse mechanism achieved in a given structure
under certain load conditions
3.3. Implementation of Mechanisms 34
each body are labeled as: 1,2,..., for body 1,2, etc. With this notation, I
12
i j
represents the i j-
components of the moment of inertia associated to mechanism "1" in body 1 and to mechanism
"2" in body 2. The rest of functions carry the same notation: U U
12...
,

d

d
12...
etc.
Additionally, if symmetry is present, some reduction process may be carried out in order to
reduce the number of equations involved [37],[51].
Consider the symmetric block of Fig. 3.3. As shown in [51], this symmetry is equivalent to

Mirror
Figure 3.3: Dynamical symmetry associated
to the sign of .
Figure 3.4: Unsymmetric non-prysmatic
rocking Block
an invariance with respect to the sign of , and, as a consequence, an associated conserved
magnitude exists (Noethers Theorem [52]).
In the most general case, however, the lack of symmetry forces to assume dierent dynamic
functions for each mechanism. This could be the case of the unsymmetric 3D non-prysmatic
block shown in Fig. 3.4. If planar motion is assumed and the mechanisms are labeled as =
(A, B), the two dierent kinetic and potential energies associated with each mechanism render:
T =

T
=A
T
A
Rotation about A
T
=B
T
B
Rotation about B
(3.51)
The same holds for the potential energy.
At this stage let s be a two dimensional vector dened as:
s = (s
1
1
, s
1
2
), (3.52)
3.4. Extension to the Riemannian Conguration Space 35
where s
1
1
and s
1
2
are dened through the sign function:
s
1
1
=
1 + sign()
2
(3.53)
s
1
2
=
1 sign()
2
(3.54)
Notice that the quantities s
1
1,2
are, in fact, the Heaviside step functions for [53].
Mechanism tensors
The procedure detailed needs not to be limited to single degree of freedom systems or, even
further, to planar mechanisms. For a single block in 3D with v edges, the expression of s
renders; s = (s
1
1
, . . . , s
1
v
), where its elements s
j
are appropriate sign functions of the generalized
coordinates involved. If only two mechanisms are possible, the expressions for s
n

render:
s
n

1 + sign(
n
)
2
1 sign(
n
)
2
(3.55)
The properties of the s symbols (Heaviside step functions) are summarized in appendix B.
For two-block assemblies, the 1-tensor (vector) s is a 2-tensor, referenced as S and given by:
S
12
= s
1
1
s
2
2
(3.56)
where the subscripts "1" and "2" belong to an index set which labels all possible mechanisms
of body 1 and 2 respectively.
In its most general form; when N blocks are considered and each body n has v(n) mechanisms:
S
123...
= s
1
1
s
2
2
s
3
3
(3.57)
where: 1 = 1, . . . , v(1) and 2 = 2, . . . , v(2), etc.
3.4 Extension to the Riemannian Conguration Space
The conguration space
It has been shown that for a given collapse mechanism, the topology of the systemof N-bodies is
equivalent to the N-torus consisting on the N-tuples of real angles (
1
,
2
, . . . ,
N
). A procedure
3.4. Extension to the Riemannian Conguration Space 36
Figure 3.5: Single block conguration space: r = ||, = arg
which has been shown to be extremely powerful consists on the complexication of the angles
[51]. Recall that for C:
S
1
S
1
= S
2
(3.58)
That is, the RM single block conguration space is a torus as shown in Fig. 3.5. Therefore, for
the general case of N bodies, the conguration manifold Mconsists of N-torus:
M = S
2
S
2
. . . S
2
= S
2N
. (3.59)
The next step is to couple the Euclidean dynamic functions with the mechanisms. This is ac-
complished through the metric tensor of the manifold.
An Hermitian inner product is assumed (see Eq. B.2 of appendix B):
< dx, dy >= g
i j
dx
i
dy
j
(3.60)
dened through the metric g
i j
:
g
i j
= S
123...
J
123...
i j
. (3.61)
3.4. Extension to the Riemannian Conguration Space 37
That is, the metric tensor carries the mechanism information and the inertia.
4
At any point p of
the manifold it is dened a tangent space MT
p
with its associated vector space.
A metric tensor eld over the manifold assigns g
i j
to each point of M. If dx
i
are the components
of the vector space of MT
p
, the line element is thus given by:
ds
2
= g
i j
< dx
i
, dx

j
> (3.63)
Kinetic energy in M
With the assumed metric, the kinetic energy results:
T =
1
2
_
ds
d
_
2
=
1
2
g
i j
<
d
i
d
,
d

j
d
> (3.64)
By using Eq. 3.61, the nal form for T is found to be:
T =
1
2
S
12...
J
12
i j

(3.65)
Potential energy in M
The potential energy in the complex manifold is simply coupled with mechanisms through a
tensor product. Since the matrix F from Eq. 3.49 depends on the mechanisms, the natural
assumption is:
U = g tr(M)S
12...
F
12...
(3.66)
4
It must be highlighted that in this expression, the metric tensor S is dened through the sign function of the
real part of the angles. That is, S
12...
= s
1
s
2
where,
s
1
1
=
1 + sign(Re[
1
])
2
(3.62)
where "Re[
1
] = r
1
cos(
1
)" is the real part of the angle
1
etc.
3.4. Extension to the Riemannian Conguration Space 38
Final expressions
First, it is noticed that the rst term of the generalized inertia tensors J (Eqs. 3.38 and 3.40)
does not depend on mechanisms. Hence, through the denitions;
I
i j

N

n=1

in

jn
I
n
(3.67)
G
i j
dA
i j
x
T
c
+ dB
i j
d
T
(3.68)
K
i j
I
i j
+ S
12...
G
12...
i j
(3.69)
Y S
12...
F
12...

2
(3.70)
Both T and U can be recast in a more condensed way as:
T =
1
2

K
T
(3.71)
and
U = g tr(M)Y. (3.72)
Finally, the resulting Lagrangian function renders:
L(

) =
1
2

K
T
g tr(M)Y (3.73)
Decomplexication of rotations
The nal step is to express the rotations in terms of real parts of the angles:
n
=
n
(r
n
,
n
). This
is accomplished by noticing that the rotation operators can be recast as:



R
r,
(3.74)
resulting in:

R
r,
= e

r
(3.75)
where the generator

has a matrix given by:
= cos() (3.76)
3.5. General Procedure to Obtain the Reduced Equations of Motion 39
As it will be shown in chapter 5, in the rocking motion limit; cos() sign()), the operator

R
r,
is unitary. This can be veried by its matrix representation in 2D.
By using Eulers formula (Eq.B.1 of appendix B):
R
r,
= cos(r)I + sin(r) =

cos(r) cos() sin(r)


cos() sin(r) cos(r)

(3.77)
Then, since cos()
2
= sign()
2
= 1, it holds; R
r,
R
T
r,
= I.
Constraint-free Equations of motion
Equations of motion can be derived through the DAlembert equations. If a Lagrangian function
is built is the usual manner as: L = T U, those equations hold:
d
d
_
L
q

L
q
= Q
q
, q = 1, . . . , N (3.78)
where "Q
q
" are the generalized forces associated to the coordinate "q". When the expressions
of U and T are substituted into L. system of Eqs. 3.78 provides the equations of motion cor-
responding to the coupled rotations of an assemblage of rigid solids. However, if additional
constraints are present, Eqs. 3.78 must be modied by the Lagrange multipliers method in or-
der to take the constraints into account.
3.5 General Procedure to Obtain the Reduced Equations of
Motion
Figure 3.6 shows through an schematic ux diagram the proposed approach to use the formula-
tion presented in this work in order to nd the probability of collapse.
1. Initiation. Collapse mechanism.
2. Derivation of geometrical parameters.
3. Implementation of constraints.
4. Construction of dynamical functions.
3.5. General Procedure to Obtain the Reduced Equations of Motion 40
Figure 3.6: Procedure for the determination of the Probability of collapse of masonry structures
Initiation of Motion. Derivation of fracture lines
The present formulation is unable to determine the collapse mechanism of a structure under
external loading. It operates once the motion is initiated by considering the collapse mechanism
as a given boundary condition. Therefore, an auxiliary tool is needed in order to complement
the dynamical approach presented herein.
Consider the case of a masonry wall subjected to an out-of-plane action. The lateral walls act
as geometric constraints and hence macro-blocks are formed. A typical collapse mechanism for
this structure is shown in Fig. 3.7 [54] consisting on ve macro-elements.
3.5. General Procedure to Obtain the Reduced Equations of Motion 41
t
st
u u
s
t
s
u
Figure 3.7: Masonry wall under out-of-plane loading. Collapse mechanism based on [54]
Derivation of geometrical parameters
From this mechanism the hinge vector

d and the density eld (r) are derived. The number
of blocks involved denes the quantities {, C, P, , R} along with the mechanisms tensors S
with their derived quantities: {, Z
c
, Q}. Then the tensors {I
n
, A, B} are found from straightfor-
ward. The next step is to derive the intermediate quantities {J, F}. Then, the coupling between
mechanisms and rotations is implemented through the tensors {K, Y} to nd U and T.
Implementation of constraints
As it is usual in Lagrangian dynamics, there are two possible approaches which can be used to
reduce the degrees of freedom:
Consideration of constraints through Lagrange multipliers.
Solve constraint equations to reduce DOF.
The dierence between these two perspectives is weather the constraints should be explicitly
considered into the equations of motion or not.
The starting conguration space is the N-torus complex manifold of Eq. 3.59. The type of
problem under study determines which approach should be used. Once the kinematic chain (i.e.
system of coupled rotating bodies) has been formed, constraints can be implemented by the
3.5. General Procedure to Obtain the Reduced Equations of Motion 42
requirement that the last axis of rotation remains constant in time. The holonomic constraints
are of the form:
g
k
(

) = 0, k = 1...m (3.79)
where m is the number of constraints. These equations reduce the degrees of freedom from N
to N m.
Lagrange multipliers are implemented into the DAlembert equations (Eq.3.78) as:
d
d
_
L
q

L
q
= Q
q
+
n
g
q
n
, q = 1, . . . , N.n = 1..., m (3.80)
The advantage of the Lagrange multipliers method is that it allows explicit monitoring of the
reaction forces ( which are in this case understood as constraint forces) [52].
On the other hand, there are situations for which the solution is better obtained by neglecting
the explicit form of the constraints and directly solving Eqs.3.79 For the particular case of 2D
systems q = 2, the above relations take a simple form:
dr
N
dt
= 0 (3.81)
which is equivalent to the condition:
d
dt
< r
N
, r
N
> (3.82)
Since r
N
=

P
N
[r], the constraint equation can be put in the form:
d
dt
[rP
n
r
T
] = 0 (3.83)
The implementation of the constraints is only achievable through symbolic manipulation since,
the explicit expression of
N
(
1
, ...,
N1
) is required.
Build-up of dynamical functions
The nal step is to build the Lagrangian function in order to derive the equations of motion.
This is accomplished also through symbolic manipulation by the programs MREDUCE.mws,
MBLOCK2D and MBLOCK3D (see annex G), based on the c Maple environment.
3.6. Summary 43
3.6 Summary
So far in this study the basic elements of the dynamics of many-block systems have been intro-
duced. Although the novel formulation presented herein may seem to have little advantage with
respect to PWT for the case of a single block, the scope of the new frame becomes evident when
the problem is translated to a computational scheme. In this case, the new approach constitutes
a good complement for other tools.
Due to the huge amount of symbolic manipulation, the reduction process and derivation of the
equations of motion is only possible if the process is aided with computer programs.
Chapter 4
Formulation for Planar Systems
In this chapter the general procedure to obtain the equations of motion for simple planar con-
gurations is presented through various cases. Solutions obtained from the computer programs
MREDUCE.mws and MBLOCK2D (see annex G), based on the procedure shown in Fig. 3.6
are shown.
4.1 General Procedure
Although there are signicant eorts to derive the equations of motion and stability conditions
for 3D rocking motion; [9],[55], planar congurations are still the most active eld of research
within the context of RM dynamics.
Part of the problem stems from the fact that the number of mechanisms rapidly increases when
3D motion is allowed. If there are N prismatic blocks involved, the number of possible mecha-
nisms for 2D motion is 2
N
whereas for 3D motion this number is 2
2N
. Furthermore, the coupling
between in-plane and out-of plane motion has shown to be a dicult problem to handle [55].
On the other hand, it has been observed from experimental tests that this 3D eects may sub-
stantially modify the dynamics (see annex F). However limitations of space and time do not
allow to take into consideration this 3D eects in the present work. Therefore, although the
formulation presented in chapter 3 is, in principle, valid for 3D congurations, this chapter is
focused on how this analytical techniques can be applied to 2D problems.
44
4.2. Single Degree of Freedom Systems 45
As it was commented in section 1.2, one of the main characteristics of stone masonry collapse
mechanisms is the division of the conguration into macro-blocks. This is due to the presence
of constraints, both geometrical and friction. The number of degrees of freedom is drastically
reduced once the collapse mechanism is formed and, hence, the overall dynamics can then be
understood as the motion of this new reduced set of macro-elements.
This fact is used by the Lagrangian formulation in order to derive the equations of motion
without considering unnecessary internal degrees of freedom. With this approach, analytical
expressions are derived, not obtainable through other techniques (DEM, FEM).
For planar systems (q = 2), the basic elements of the dynamics introduced in chapter 3 take
simple expressions which allow great simplication and reduction.
4.2 Single Degree of Freedom Systems
The most representative single degree of freedom (SDOF) structures concerning RM dynamics
are:
Single-block structures.
Closed-Chain Structures.
Four-hinged masonry arch.
Part III of this work deals with single block structures in detail. There, stability and dynamics
are extensively studied.
On the other hand, some apparently complex congurations, turn out to be simple from the
dynamical point viewsince only one degree of freedomis involved. This is the case of the planar
closed-kinematic chain structures where, due to the constraints, the motion can be understood
as a SDOF problem. In this sense, the stability of a four-hinged masonry arch can be understood
in terms of the single block dynamics.
4.2. Single Degree of Freedom Systems 46
Closed-Chain Structures
The so called closed kinematic chain structures arise in many problems when the overall be-
havior of stories in a building or temples. Figure 4.1 shows the main parameters of a closed
kinematic chain structure. Since there are three bodies involved and q = 2;

= (
1
,
2
,
3
).
Figure 4.1: Closed-kinematic chain structure. Denition of parameters
However, the last hinge remains xed during the motion, which materializes a holonomic con-
straint. The requirement of the last hinge xed implies two relations of the form of Eq. 3.79.
As a consequence, the reduction process leads to a DOF reduction for 3 to 1:
{
1
,
2
,
3
}

2
=
2
(
1
)

3
=
3
(
1
)
(4.1)
The program MREDUCE.mws is used to obtain the nal degree of freedom as function of
1
and
2
from Eq.3.83. The resulting equations are:
b(
2
+
3
) = 0
(L + a)
1
+ L
2
a
3
= 0 (4.2)
where "L", "a" and "b" are geometrical parameters dened in Fig.4.1. The solution is:

2
=
1

3
=
1
(4.3)
4.2. Single Degree of Freedom Systems 47
Figure 4.2: Closed-kinematic chain structure. Solution after DOF reduction
Figure 4.2 presents the motion displayed after the DOF reduction with the aid of the program
MBLOCK2D.mws. The dynamical problemof the closed kinematic chain has then be translated
into an equivalent rocking block system.
Four-hinged Masonry Arch
The four-hinged mechanism of a masonry arch under lateral loading is another SDOF problem
with special interest in masonry construction (see Heyman [11]). In fact, this mechanism can
be considered as a special type of a closed-kinematic chain problem. The problem is well for-
mulated when the geometry, friction angle and external action are provided. Based on these
parameters, the algorithm computes the load factor, line of thrust and the associated collapse
mechanism. For circular arches (see Fig.4.3), the geometry is specied by providing the param-
eters.
1. Number of dowels (ND).
2. Span (L).
3. Buttress width (2H).
4. Buttress height (2B).
4.2. Single Degree of Freedom Systems 48
Figure 4.3: Line of thrust of a masonry arch under lateral load proportional to its own weight.
Solution computed with ARCOTSAM.mws
The external action is modeled as horizontal forces concentrated at the centers of mass of each
body, with a magnitude proportional to the body mass. Finally, the friction angle is assumed
in the range 30 35
o
.
Finally, the line of thrust, shown in Fig.4.3, is computed with the program ARCOTSAM.mws
within the c Maple environment (see appendix G). The collapse mechanism consists on the
formation of four plastic hinges. The four hinge is likely to appear at the buttress, as shown
in Fig.4.4 This mechanism was obtained for the parameters: (ND = 12, L = 1.5m, 2H =
Figure 4.4: Collapse mechanism I of a four-hinged masonry arch. Solution after DOF reduction
2.2m, 2B = 0.6m, = 30
o
). However, variation in the values of the parameters can lead to
another collapse mechanism. Figure 4.5 shows the second most probable collapse mechanism.
In this case one of the buttress remains xed and the four hinge appears within the dowels.
4.3. Multiple Degree of Freedom Systems 49
Figure 4.5: Collapse mechanism II of a four-hinged masonry arch. Solution after DOF reduc-
tion
Once the collapse mechanism is given, the program MREDUCE.mws computes the geometrical
properties based on the equivalent SDOF system. Then an equivalent Lagrangian function is
derived through the procedure described in section 3.5. Then, the dynamical problem can be
addressed by the methodologies of single block dynamics.
4.3 Multiple Degree of Freedom Systems
Ranging from 1 to N stacked blocks, both experimental and analytical contributions arose in the
context of masonry dynamics in the last decades. However, when multiple degrees of freedom
are considered, the amount of works using Housner-type equations decreases dramatically.
On the other hand, on the basis of the formulation of multi body dynamics with deformable
contacts, the number of contributions is huge. In this regard, special mention deserve those
works using DEM methods in their approach.
Two stacked blocks.
Column formed by N stacked blocks.
De Felice [36] has used a DEM code to estimate the fragility of masonry walls subjected to
out-of-plane loading through a simplied 2D models.
4.3. Multiple Degree of Freedom Systems 50
Two-block assemblies
Two block assemblies congurations have been studied both analytically [9], [33], [56], and ex-
perimentally [9]. The study of two-block assemblies has been applied to the seismic assessment
of crane bridge-like structures in [57]. On the other hand, the dynamics of two macro-elements
has been applied by Doherty et al. ([35]) and Grith et al. [58]) to the assessment of the out-
of-plane stability of unreinforced masonry walls subjected to seismic action. In these works
the constraints imposed lead to the idealization of the model into a single degree of freedom
structure.
The basic elements of the planar dynamics for

= (
1
,
2
) take very simple expressions:
1. The matrices are equal to i
2
, where
2
is one of the Pauli matrices (see annex B).
2. Rotation matrices take the form
R =

1
O
2x2
O
2x2
R

(4.4)
where R

1
and R

2
are planar rotations with matrix R

k
= exp(i
2

k
), and O
2x2
, I
2x2
repre-
sent the null and identity 2x2 matrices respectively.
3. The chain operator matrix is:
C =

O
2x2
I
2x2
O
2x2
O
2x2

(4.5)
4. There are four possible mechanisms for two coupled blocks (see Fig. 4.6). Table 4
summarizes the "s" symbols and the associated S
1,2
mechanism tensors for this case:
Therefore, the mechanism tensor is simply the 4x4 identity matrix.
With these reduced elements, the dynamical functions are easily obtained. It has been stated
that the resulting Lagrangian function can be understood as the Lagrangian of a double inverted
pendulum system. Figure 4.6 plots the solutions obtained from the program MREDUCE.mws.
The cases
2
= 0 are not considered as a two-block mechanism since they are a SDOF.
4.3. Multiple Degree of Freedom Systems 51
Mech s
2
1
s
1
2
s
1
2
s
2
2
S
11
S
12
S
21
S
22
(+,+) 1 0 1 0 1 0 0 0
(+,-) 1 0 0 1 0 1 0 0
(-,+) 0 1 1 0 0 0 1 0
(-,-) 0 1 0 1 0 0 0 0
(+, +) (, )
(+, ) (, +)
Figure 4.6: Collapse mechanism of two rigid stacked blocks
Column
Systems with N stacked rocking blocks forming a column have been studied by Augusti [17],
Sinopoli [34]. More recently, Psycharis [59] has used DEM code to investigate the stability if
classical columns under harmonic and earthquake excitations.
There are NDOF for planar motion corresponding to the coupled rotations of

= (
1
,
2
, ...,
N
).
If all the blocks are equally sized with dimensions: height=2h and width=2b, the expressions
of the d and x
c
vectors for each mechanism is very simple. Table 4.1 presents the d values for
each mechanism.
4.3. Multiple Degree of Freedom Systems 52
n/n + 1 1 2
1 (0,2h) (2b,2h)
2 (-2b,2h) (0,2h)
Table 4.1: Hinge vectors for a system of N stacked blocks
On the other hand, the vectors x
c
only depend on the hinge n of body n. Their expression is
also very simple:
x
n
c
=

(b, h); n = 1
(b, h); n = 2
(4.6)
Figures 4.7 present a column formed by 14 equal stacked prismatic blocks. In Fig 4.8 the
Figure 4.7: Column formed by 14 equal stacked prismatic blocks
new conguration of points has been computed by solving Eqs. 3.78 for an equally distributed
angular rotation of = 15
o
. The solution is displayed through the program MBLOCK2D.mws.
4.4. Summary 53
Figure 4.8: Planar motion of of a column formed by 14 stacked prismatic blocks
4.4 Summary
This section presented a possible approach to address the motion of rigid block assemblages
without considering interface equations -such as joint springs or dampers- from a novel point
of view. The main results are summarized in the following.
Complex structures such as four-hinged masonry arch can be reduced to dynamical sys-
tems with a single degree of freedom. The Lagrangian formulation uses this DOF reduc-
tion in order to derive the equations of motion analytically.
Planar motions of stacked blocks are characterized by very simple expressions of the d
and x
c
vectors due to symmetry. For these cases, constraints do not allow DOF reduction.
The formulation does not intend to be a substitute for the existing standard formulations (DEM,
FEM, etc.) but to provide an alternative tool for the blocky masonry modeling. Future research
is needed in order to address both experimentally and numerically the probability of collapse
for the simplied models presented herein.
Part III
Single Block Dynamics and Stability
54
Chapter 5
Complex Formulation for Single Block
Systems
In this chapter the equations of motion of a single block under horizontal harmonic forcing are
derived from the results of the formulation of multiple-body assemblies described in chapter 3.
The so-called Rocking Motion limit is introduced as a working hypothesis, which is shown to
be valid if one of the parameters of the model tends to zero. Two mechanical analogies are
included to illustrate the physics of the problem. Finally, a comparison with the classical theory
is given.
5.1 The RM limit
Let the angle be reformulated as a complex quantity by the procedure described in section 3.4
of chapter 3:
= re
i
(5.1)
The following additional hypothesis are now introduced:
RM limit

cos() sign()
sin() 0
(5.2)
55
5.2. Free Rocking Motion 56
This is equivalent to consider that under such conditions the real part of equals or, more
precisely, that the RM limit implies a decomplexication of the problem.
5.2 Free Rocking Motion
The Kinetic energy K and the potential energy U can be written as,
T =
1
2

K
T
(5.3)
U = g tr(M)Y (5.4)
see Eqs. 3.71 and 3.71, which, for a single rocking block, render:
tr(M) = m (5.5)
C = O (5.6)
Z
c
= x
c
(5.7)
A = B = O (5.8)
S

= s

(5.9)

2
= e
2
=

0
1

(5.10)
where m equals the mass of the block, x
c
is the matrix vector of the center of mass of the block
in a reference frame with origin at one of the block corners O or O

(see Fig 2.1 of chapter 2),


and s

equals the signum tensors of Eqs. 3.54 dened in chapter 3.


Potential energy
Since C = O, the expression of Y, with the above conditions renders:
Y = s

c
Re
2
(5.11)
5.2. Free Rocking Motion 57
By using the matrix of the decomplexied rotation operator of Eq. 3.77, the potential energy
renders:
U = gm[sin(r)(cos()s
1
x
1
c,1
+ cos()s
2
x
2
c,1
) + cos(r)(cos()s
1
x
1
c,2
+ cos()s
2
x
2
c,2
)] (5.12)
where by x
1
c,1
it is meant the rst (horizontal) component of the center of mass vector corre-
sponding to mechanism 1, etc. From Figure 1.5 it is clear that:
x
1
c
= (b, h)
x
2
c
= (b, h) (5.13)
(5.14)
and by using the properties of the s symbols (Eqs. B.18), Eq. 5.12 renders:
U = gm[sin(r)b + cos(r)h] (5.15)
where the rocking limit has been used (sign() cos()).
In terms of the semi-diagonal of the block and the angle , dened in Fig. 1.5 it is found that
the potential energy depends solely on the absolute value of the angle and reads
U = MgR[cos( r) cos()] (5.16)
Kinetic energy
For this case, the vectors are simply given by: = . Taking into account that matrices A and
B are zero, J is simply equal to the moment of inertia for each mechanism. Therefore K = s

.
Since the moments of inertia with respect to O and O

are equal; I
1
= I
2
= I, the kinetic energy
reads:
T =
I
2

(5.17)
Derivation of the dynamics
The Lagrangian function is built in the usual way as L
0
= T U, or
L
0
=
I
2
(r
2
+ r
2

2
) MgRcos( r) (5.18)
5.2. Free Rocking Motion 58
The mechanical energy of the problem E reads
E =
I
2
(r
2
+ r
2

2
) + MgRcos( r) (5.19)
Since the Lagrangian function does not depend explicitly on , the canonical conjugate mo-
mentum, p

, is a well-dened constant quantity:


p

=
L
0

= Ir
2

(5.20)
The notation is now changed by introducing, l
0
: p

l
0
. In its present form, the problem is
equivalent to that of a single particle with kinetic energy K = (I/2)r
2
moving in an eective
potential V
e f f
given by:
V
e f f
(r) =
l
2
0
2Ir
2
+ MgRcos( r) (5.21)
This potential adds an innite repulsive barrier to U(r), at r = 0. For a given value of energy
E the position of the corresponding turning points r
min
and r
max
can be obtained by numerically
solving the equation E = V
e f f
(r). However, for small l
0
the following (approximate) analytical
solution can be derived for r
min
,
r
min
=
l
0

2I(E U
0
)
(5.22)
where U
0
U(r = 0) = MgRcos().
It should be pointed out that this reduction process is equivalent to the introduction of the
Routhian function [52]:
R L p

(5.23)
which eliminates the kinhostenic variable . The Routhian is equivalent to a modied La-
grangian whose potential is V
e f f
.
5.3. Implementation of Impact and Seismic Actions 59
5.3 Implementation of Impact and Seismic Actions
Seismic and damping actions are now introduced via DAlembert equations (Eqs. 3.78), for
both seismic and damping actions:
d
d
_
L
0
r

L
0
r
= Q
d
r
+ Q
e
r
(5.24)
d
d
_
L
0

L
0

= Q
d

+ Q
e

(5.25)
Here Q
d
r
and Q
e
r
represent the generalized damping and earthquake forces with respect to the
canonical variable r respectively. The same holds for Q
d

and Q
e

for .
Impact model. Phase dynamics
Damping eects do not depend on the sign of , that is, the impact force is the same indepen-
dently of which mechanism is taking place. Therefore, it is assumed that
Q
d

= 0 (5.26)
Furthermore, the impulsive force is considered vertical and acting only through the edges O
and O

just at the moment of impact and with a magnitude equal to a half of the block weight
W = Mg (see Fig. 5.1).
If the reaction force through an edge is referenced as F, it must be noticed that this force changes
instantaneously in an amount of W/2 during the impact. This suggests to consider F of the form:
F =
Mg
2
F

(5.27)
being F

a Dirac-delta force. The implementation of such an impulsive force has yield excellent
results in a previous work [50].
On the other hand, since the generalized force with respect to (or to its absolute value r) is
equal to the moment of the reaction force F measured from the centers of rotation O and O,
the expression for Q
d
r
is simply: Q
d
r
= 2bF, or, equivalently:
5.3. Implementation of Impact and Seismic Actions 60
Figure 5.1: Impulsive forces during impacts
Q
d
r
= MgRsin()F

(5.28)
In order to obtain the mathematical form of F

, two issues should be taken into account. First,


is associated to the sign of and, accordingly,

is related to the change of the sign of


with time. Second, the eect of the impact forces is to reduce the absolute value of the angular
velocity when the change of sign takes place.
The simplest form for a generalized coupling between r

and the change of sign is:


F

= c

(5.29)
where "c" is a constant.
Evaluation of the constant "c"
To obtain the value of "c" just introduced, the power dissipated by Q
d
r
is considered. Taking into
account Eqs. 2.5, 5.28 and 5.29 and recalling that the power P, dissipated by a damping force is
equal to the product of that force by the velocity: P dE/d = Q
d
r
r

dE
d
= r

Q
d
r
= I p
2
c(r

)
2
sin()

(5.30)
5.3. Implementation of Impact and Seismic Actions 61
From Eqs. 5.19- 5.22 it holds that: (r

)
2
= 2(E V
e f f
)/I and this can be substituted in Eq. 5.30
to give:
dE
E V
e f f
(r)
= 2p
2
c sin()d (5.31)
This expression must be integrated during an impact. The left-hand-side is integrated in energies
between E
b
and E
a
(two consecutive impacts), while the right-hand-side is integrated in the
phase .
Since, at impact, the contribution to the change of energy, E = K+V
e f f
, due to V
e f f
is negligible
compared to that corresponding to the kinetic energy, dV
e f f
can be neglected in the numerator
of Eq. 5.31
dE
E V
e f f
(r)
=
d(K + V
e f f
)
K

dK
K
(5.32)
leading after integration to
ln
_
K
a
K
b
_
= 2p
2
csin() (5.33)
Because, it also holds (see appendix, Eq. C.15) that for l
0
0, = , Eqs. 5.33 and 2.13 lead
to
c =
ln()
p
2
sin()
(5.34)
Thus, the form of the generalized damping force results in:
Q
d
r
= (
ln()I

(5.35)
Implementation of harmonic forcing
When a horizontal force is applied to the block, the mirror symmetry does not hold any more.
The dynamics are no longer invariant under the change of sign of , and, accordingly, l
0
be-
comes, in general, a function of time. However, for l
0
suciently small, it will be shown that
this variation is negligible.
The eect of the acceleration considered in Eq. 2.4 consists on a horizontal force of magnitude
Ma(), applied at the center of mass of the block.

F = Ma() (5.36)
5.3. Implementation of Impact and Seismic Actions 62
As a consequence, the generalized forces corresponding to the variables r and are
Q
e
r
= Ma()
X
cm
r
(5.37)
Q
e

= Ma()
X
cm

(5.38)
being X
cm
the horizontal coordinate of the center of mass measured with respect to the center of
rotation:
X
cm
= sign()Rsin( ||) (5.39)
As shown in the appendix (see Eq. C.17 and Fig. C.2), for l
0
small, the sign of can be given
by cos(). Therefore, Eq. 5.39 renders
X
cm
= Rcos() sin( r) (5.40)
which, once substituted in Eqs. 5.37 and 5.38, leads to
Q
e
r
= Ma()Rcos( r) cos()
Q
e

= Ma()Rsin( r) sin()
(5.41)
In the limit l
0
0, only takes the values n (n=0,1,2,...) and Q
e

vanishes. Thus, it has been


shown that the hypothesis introduced in section 5.1 are justied if l
0
0.
Now, the non-dimensional variable , equal to |x| and given by: (r/) is introduced. In
addition, the Lagrangian function L
0
is normalized by the denition: L L
0
/(I p
2

2
). The new
Lagrangian results in
L =
1
2
(
2
+
2

2
)
1

2
cos( ). (5.42)
The corresponding normalized DAlembert forces with respect to , considering Eqs. 2.4, 5.35
and 5.40 render
Q
d

= (I p
2
)
_
ln()

_

(5.43)
Q
e

= (I p
2
) cos(t + ) cos( ) cos() (5.44)
Taking into account that;
L
0
r
= (I p
2
)
L

(5.45)
5.4. Mechanical Analogies 63
d
d
_
L
0
r

_
= (I p
2
)
d
dt
_
L

_
(5.46)
DAlembert equations lead to:
=

sin( ) +
ln()

+ cos(t + ) cos( ) cos() (5.47)


and
d
dt
_
L

_
= 0 (5.48)
This last equation simply expresses the fact that the quantity l, dened by l
2

is a constant
of the motion. In fact, l is the non-dimensional analogous of l
0
; l = l
0
/(I p
2
).
An autonomous system of dierential equations can be formed by embedding the space of
variables in a higher dimension one, dening:
(
1
,
2
,
3
,
4
) (, , , t) (5.49)
Due to the eect of a nite interaction, the integrator will take certain intermediate values for
the velocity at the impact time. This average results in adding a factor of 1/2 (for a more detailed
discussion, see [50]). Therefore, a factor of 2 should be included in the impulsive term in order
to make theoretical and numerical results compatible.
The nal system of equations (where the earthquake phase assumed to be zero) is

1
=
2

2
=
l
2

3
1

sin(
1
)

+
2 ln()
2
l

2
1
+ cos(
4
) cos(
3
) cos(
1
)

3
=
l

2
1

4
=
(5.50)
5.4 Mechanical Analogies
Through simple transformations, two relevant mechanical analogies can be derived from the
complex formulation (CF). On one hand, the problem is reset to be equivalent to an inverted
pendulum. Then, the term inverted pendulum structures, so extensively used, gets its full mean-
ing.
5.4. Mechanical Analogies 64
The fact that the dynamics can be recast in the form of a pendulum opens the possibility of
a generalization of N rocking blocks (as in a column) to N-inverted pendula. For this case a
signicant amount of research is found in literature.
On the other hand, it is also reset how CF can be recast in a two-body central problem in the
complex plane.
The inverted pendulum Hamiltonian
From the non-dimensional Lagrangian of Eq. 5.42, momenta are given by:
p


L

, (5.51)
p

, (5.52)
where p

= l = p

/(I p
2
). The Hamiltonian is dened as usual
H

j
p
j
q
j
L, (5.53)
or
H =
p
2

2
+
p
2

2
2
+
1

2
cos[(1 )]. (5.54)
Using this Hamiltonian, a canonical point transformation is performed from (, , p

, p

) to
(Q

, Q

, P

, P

), with the aid of the generating function


F(, , P

, P) = (1 a(1 ))P

+ P

(5.55)
where the non-dimensional quantity a / has been introduced. The new variables are then
dened through the equations
p

= aP
x
, (5.56)
Q


F
P

= (1 a(1 )), (5.57)


p

= P

, (5.58)
Q


F
P

= , (5.59)
5.4. Mechanical Analogies 65
and the new Hamiltonian in variables (Q

, P

, P

) (up to a constant term) reads


H =
P
2

2
bcos(Q

) +
P
2

2(Q

c)
2
(5.60)
where the parameters b and c are given by b
1
(a)
4
and c (1 a), respectively.
By use of Eq. 5.58, recalling that p

= l, and by labeling the canonical variables (Q

, P

)
as (q, p), this Hamiltonian renders:
H(p, q) =
H
P
,..,
p
2
2
b cos(q) +
Barrier
,..,

2
2(q c)
2
. (5.61)
If l is set equal to zero, the resulting Hamiltonian H
P
corresponds to an inverted pendulum. It
is possible then to understand RM as an inverted pendulum with an innite potential barrier at
q = c. Thus, the denition given by Housner [7] as inverted pendulum structures nds its full
meaning.
The eective potential of this Hamiltonian is given by
V
e
(q) =
l
2
2(q c)
2
b cos(q). (5.62)
Figure 5.2 depicts the typical form of the well for a = 0.245 and l = 0.01. It is clear that for
q and free RM the motion is unbounded. It must be noticed that due to the singularity at
q = c, if the parameter l was equated to zero, the potential barrier would be removed and then,
the oscillatory motion would be destroyed.
Associated two-body problem for slender blocks
For slender blocks, the trigonometric term in Eq. 5.62 can be expanded as a Taylor series in
and . Considering up to second order in the expansion
cos((1 )) 1

2
2
(1 )
2
, (5.63)
and neglecting the superuous additive constant, the Hamiltonian of Eq.5.54 can be recast as
H =
p
2

2
+
p
2

2
2
+

2
2
. (5.64)
5.4. Mechanical Analogies 66
c
b
Ve(q)
q
0
Figure 5.2: Inverted Pendulum potential. a = 0.245 and l = 0.01
This is the Hamiltonian of a two-body central problem in the complex plane with the potential
v() =

2
2
. In the absence of any forces other than gravity, the system is conservative, and
the mechanical energy E given by
E =

2
2
+

2
2
+

2
2
, (5.65)
is an integral of the motion.
The eective potential
V
e f f
() =
l
2
2
2
+

2
2
(5.66)
has the form shown in Fig. 5.3. Notice that V
e f f
() is the non-dimensional analogy of V
e f f
(r).
For 0 it behaves as an innite barrier, whereas for > 0 it has the shape of a linear term
plus an inverted parabola.
Let
min
and
max
be the points where the inversion of the motion takes place. They can be
obtained by solving the equation E = V
e f f
():

4
2
3
+ 2E
2
l
2
= 0 (5.67)
In general,
min
and
max
are complicated functions of E and l. Nevertheless, for l close to zero,
good eective approximations for these points can be obtained. For
min
the solution is very
5.5. Comparison with Housner Theory 67
Xmin Xmax
E
Veff(Xc)
x
0
Veff(x)
Xc
Figure 5.3: Eective potential V
e f f
for l = 0.01 and minimum at
c
= l
2/3
= 0.0464. A bounded
state with its corresponding turning points is also shown.
close to zero and then, the third and fourth order in in Eq. 5.67 can be neglected.
On the other hand, for
max
, far away from the barrier at = 0, the centrifugal term in the
potential can be neglected. within these limits, the respective values are:

min

l

2E
(5.68)

max
1

1 2E (5.69)
As expected, the value for
min
coincides with the impact parameter in the typical scattering
problem by a central eld.
5.5 Comparison with Housner Theory
For small x and , analytical solutions do exist for Eqs. 2.6 (see, for instance [15]). Therefore,
it is possible to compare the CF method of Eqs. 5.50 with analytical expressions.
Eqs. 5.50 have been integrated numerically using a gear method (see appendix G). In this
process the following should be taken into account:
5.5. Comparison with Housner Theory 68
The integration step must always be smaller than l. In the computer implementation, a
value of l/10 has been adopted as default.
The value of r (or its non-dimensional analogy ) is lower-bounded by the value r
c
(or

c
) which corresponds to a minimum of the eective potential V
e f f
(r), or V
e f f
().
The value for r
c
or
c
can be obtained by solving:
dV
e f f
(r)
dr
r=r
c
= 0 (5.70)
which leads to the nonlinear equation
MgRsin( r
c
) =
l
2
0
Ir
3
c
(5.71)
In terms of l and , this equation leads to
sin( )
3
c
= l
2
(5.72)
For small values of and
c
, the following approximate value can be obtained

c
l
2/3
(5.73)
Therefore, a reasonable condition for numerical integration of the system of Eqs. 5.50 is

1
l
2/3
.
Next, two dierent examples are used to validate the proposed novel formulations. The ob-
tained results plot the real part of the non-dimensional angle x versus the non-dimensional time
t, that is (t, Re(x) = cos())), or equivalent, (t,
1
cos(
3
)).
As a rst example, the model for dierent values of l will be compared with Housner the-
ory. Under these conditions, the performance of the damping term Q
d

is evaluated. Fig. 5.4


illustrates the results, being clear that convergence to the theoretical solution is obtained by in-
creasingly lowering the value of l. The parameters used to obtain the results of Fig. 5.4 were:
= tg
1
(1/4), = 0.925 and initial conditions; (
1
(0) = 0.5,
2
(0) = 0.0,
3
(0) = 0.0,

4
(0) = 0.0).
5.5. Comparison with Housner Theory 69
Figure 5.4: Comparison between Housner theory and complex formulation for dierent values
of l under free rocking motion.
Figure 5.5: Comparison between Housner theory and complex formulation for l = 10
5
.
5.6. Summary 70
The next example validates Q
d

and an "accurate" value of l = 10


5
was chosen for the calcu-
lation. Fig. 5.5 illustrates the perfect agreement between the theoretical solution and the novel
complex formulation. The parameters used now are: = tg
1
(1/4), = 0.925, and initial
conditions; (
1
(0) = 0.061,
2
(0) = 0.061,
3
(0) = 0.0,
4
(0) = 0.0).
5.6 Summary
So far, some of the theoretical development has been reported. The main results obtained
through the complex formulation of single block dynamics are:
1. Complex formulation allows the use of symmetry in DOF reduction.
2. The RM Hamiltonian can be canonically transformed into an inverted pendulum Hamil-
tonian.
3. Single block RM dynamics can be understood as a two-body central problem through
appropriate canonical transformations.
4. Agreement with Housner piecewise theory is excellent both for free and harmonic motion
regimes.
In the next chapter, the results from an experimental research carried out at the National Labo-
ratory of Civil Engineering of Portugal will be described. From those results, the validation of
the new theoretical approaches introduced will be further accomplished.
Chapter 6
Analysis of the Response
This chapter presents the results obtained from the experimental research carried out at the seis-
mic table of the National Laboratory of Civil Engineering (LNEC) (a more detailed description
of the experiments can be found in appendix E). Analysis of the measured response and compar-
ison with theory are included. Additionally, a numerical simulation has been reported in order
to complement those orbits which, although considered as fundamental, could not be found by
experimental tests.
6.1 Description of the Experimental Research
A set of experiments at the LNEC seismic table (see appendix E) on four blue granite specimens
were conducted. The experimental program consisted on three types of tests:
free rocking motion,
harmonic motion,
random motion.
First, the free RM tests allowed the identication and calibration of the main parameters present
in the RM dynamics. Then the condition of the initiation of rocking:
PGA a
w
(6.1)
71
6.1. Description of the Experimental Research 72
where PGA is the peak ground acceleration and:
a
w
g tan() (6.2)
is the so-called "West" acceleration, which allowed the set-up of the operating range in ampli-
tudes of the seismic table. Subsequent sine-sweep tests were carried out in order to nd out
approximately the principal resonance "frequencies" of the system which would x the experi-
mental functional range for the seismic table.
Figure 6.1 plots the non-dimensional energies of the block as function of frequency for a run-up
(and run-down) test on one specimen. The frequency was smoothly increased (and decreased)
0
0.01
0.02
0 1 2 3 4 5 6
E
n
e
r
g
y

/

I
p
2
Freq.(Hz)
R-UP
R-DOWN
Figure 6.1: Experimental energies for sine-sweep tests on specimen 1- Maximum table dis-
placement: 5mm. Frequency range (0.5Hz 5.0Hz)
for a constant amplitude of maximum table displacement of 5mm. It is clear that the range of
frequencies is between 3Hz and 5Hz. Similar tests were carried out for all specimens.
Subsequent harmonic tests evidenced the basic features of the dynamic behavior of single blocks
undergoing RM regime. Many of the results found had already been reported in other experi-
mental works ([9],[14],[39]).
6.1. Description of the Experimental Research 73
On the other hand, the behavior of the RB under earthquake conditions is examined through
random tests. The Specimens were subjected to random motion in order to study their dynami-
cal behavior under earthquake motion. Thirty synthetic earthquakes, compatible with the design
spectrum proposed by Eurocode 8, were generated. In order to identify them, they were labeled
with integer numbers as they were generated.
The "platform" of the spectrum is located between 0.1 and 0.3 seconds, with a spectral acceler-
ation of 7 m/s
2
and PGA = 2.8 m/s
2
. The number and type of tests carried out is summarized
in Table 6.1.
Table 6.1: Types and number of tests
Specimen Free Harmonic Random Total
1 3 19 22
2 7 39 44 90
3 3 11 103 117
4 5 16 31 52
Total 18 85 178 281
Characteristics of testing specimens
The experimental tests were carried out on four granite stones specically manufactured for the
tests (Figure 6.2). The mechanical properties of the granite are given in Table 6.2 [60]. Each
stone has dierent geometrical dimensions (Figure 6.2, Table 6.3). The specimens 1, 2 and
3 were dimensioned in order to have a ratio Height-Width (h/b) of 4, 6 and 8, respectively,
keeping the height constant. In order to minimize the three-dimensional (3D) eects, the blocks
have a ratio thickness-width (d/b) of 3.
In order to delay as much as possible the continuous degradation at the block corners occurring
at every impact, the stones were manufactured with a small cut of 45
o
at their bases, with
approximately 5mm side.
6.1. Description of the Experimental Research 74
Figure 6.2: Specimens for experimental tests
Table 6.2: Mechanical properties of granite stone
Density 2670 Kg/cm
3
Axial Compression Strength 134 MPa
Bending Strength 14.7 MPa
Apparent Porosity 0.4 %
Water Absorption 0.2 %
Abrasion Resistance 1.2 mm
Shock Resistance 30 mm
Specimen number 4 was specically designed with a large 45
o
cut (40 mm) in order to compare
its performance with the rest of the stones. In addition, a specimen of the same material was
used as the base where the blocks rock. This foundation was anchored to the seismic table by
means of four iron bolts.
6.1. Description of the Experimental Research 75
Figure 6.3: Geometry of the blocks tested. Stones 1,2,3 (Left) and 4 (Right)
Table 6.3: Stone dimensions.
Stone Width 2b (em) Height 2h (m) Thickness 2d (m)
1 0.25 1.000 0.754
2 0.17 1.000 0.502
3 0.12 1.000 0.375
4 0.16 0.457 0.750
Base 1.00 0.250 0.750
Relevant parameters of the RM dynamics
It has been observed along successive experimental investigations [9],[14],[16],[39],[40], that
the set of parameters {, p, } introduced in chapter 2 plays a fundamental role for the planar
Rocking Motion problem. These quantities depend solely on the geometry of the specimen:
= tan
1
_
b
h
_
; p =
_
3g
4R
; = 1
3
2
sin
2
() (6.3)
6.1. Description of the Experimental Research 76
Some authors [9],[15],[16],[32], [44], [61], [62], studied the dynamics of the RB under dierent
conditions. From those studies it is possible to conclude that:
1. the theoretical parameters are dierent from those obtained by laboratory tests,
2. the friction between the foundation and the block is not innite and, thus, it is possible to
have sliding behavior,
3. the response of a RB is very sensible to the boundary conditions, the impact (coecient
of restitution) and the type of the base motion (harmonic, random, frequency contents,
etc).
These evidences have also been found during the experimental tests reported here, along with
other eects such as 3D rotation and degradation of specimens.
The parameters and can be obtained by Eq. 6.3 or, alternatively, by means of free rocking
motion test (see appendix E for details). Apossible explanation of the disagreement between the
theoretical and experimental values of the parameters (see Table E.1) is that the blocks do not
fulll completely the hypotheses proposed by the classical theory (section 2.1). Lipscombe [9]
reported a detailed study on the possible causes of these variations, such as; bouncing eects,
exural deformations, 3D behavior, etc. The present work does not include a study accounting
for such disagreement between theoretical and test parameters. The more straightforward tech-
nique of tting the parameter values to those yielding a best match with experiment is followed
here.
The reason for this choice relies upon the fact that those parameters (as the whole RM problem
itself) are considered as intrinsically statistical. The main eort in this work is to establish
a theoretical framework on which that statistical research could be developed. Therefore, the
aim is to nd out patterns and behavior concerning stability criteria along the whole parameter
range, as it will be accomplished in chapter 7.
6.2. Types of Response and Their Spectra 77
6.2 Types of Response and Their Spectra
Free RM and the Housner frequency
As indicated above, free RM tests were used to calibrate the model in order to nd the values
of the parameters (, p, ).
Denition of Housner frequency
From Eq. 2.12 it is obtained the so-called "Housner frequency", f
H
, can be dened as
f
H
=
p
4 cosh
1
(


0
)
(6.4)
This equation makes sense only in the free RM regime with zero initial velocity and an initial
angle equal to
0
. Eq.6.4 is not in general valid for forced motion. However, as it will be
shown, interesting results are found under forced motion regimes when it is used as a reference
frequency.
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0 0.5 1 1.5 2 2.5 3 3.5 4
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
H
f
max
a) (0) = 3.5
o
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.18
0 0.5 1 1.5 2 2.5 3 3.5 4
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
H
f
max
b) (0) = 2.6
o
Figure 6.4: Specimen 1. Fourier spectrum of free RM with initial amplitude; a) x
0
= 3.5
o
, b)
x = 2.6
o
. Main frequency and Housner frequency computed with Eq. 6.4 are also shown.
Figures 6.4 and 6.5 show two spectral analysis on experimental responses of specimens 1 and 2
for free RM regime and dierent initial conditions. The maximum peak frequency along with
6.2. Types of Response and Their Spectra 78
the frequency computed from equation 6.4 are also shown. Similar results are obtained for the
rest of specimens. Two relevant facts are to be highlighted. First, the Housner frequency lies
0
0.05
0.1
0.15
0.2
0.25
0.3
0 0.5 1 1.5 2 2.5 3
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
H
f
max
a) (0) = 4.11
o
0
0.1
0.2
0.3
0.4
0.5
0.6
0 0.5 1 1.5 2 2.5 3
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
H
f
max
b) (0) = 4.28
o
Figure 6.5: Specimen 2. Fourier spectrum of free RM with initial amplitude; a) x
0
= 4.11
o
, b)
x = 4.28
o
. Main frequency and Housner frequency computed with Eq. 6.4 are also shown.
close to the peak and, second, besides de main frequency, there is a cascade of frequency peaks
with decreasing magnitude as they move away from the main frequency.
Comparison between theory and experiment
Figures 6.6 and 6.7 report the comparison between theory and experiments for all the specimens.
Inadequate agreement was observed when the solution obtained from numerical integration of
Eqs. 5.50 was tested using the theoretical geometrical data. However, good agreement between
theory and experiment is achieved when the tted values (table E.1) are used.
6.2. Types of Response and Their Spectra 79
-3
-2
-1
0
1
2
3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

(
o
)
(s)
Theory
Experiment
a) Specimen 1
-4
-3
-2
-1
0
1
2
3
4
0 1 2 3 4 5 6 7

(
o
)
(s)
Theory
Experiment
b) Specimen 2
Figure 6.6: Comparison theory-experiment for free RM with tted parameters of Table E.1; a)
specimen 1, b) specimen 2.
-4
-2
0
2
4
0 2 4 6 8 10 12 14

(
o
)
(s)
Theory
Experiment
a) Specimen 3
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
0 0.5 1 1.5 2 2.5 3 3.5

(
o
)
(s)
Theory
Experiment
b) Specimen 4
Figure 6.7: Comparison theory-experiment for free RM with tted parameters of table E.1; a)
specimen 3, b) specimen 4.
6.2. Types of Response and Their Spectra 80
Harmonic loading. Constant amplitude
Classication and main parts of response
In the following, the Spanos and Koh [14] classication of periodic responses is adopted. Orbits
are labeled by pairs of positive integers (m, n), where 2m represents the number of impacts per
interval of repetition whereas n is the number of periods of the external force in that interval.
Additionally, modes are called Symmetric or Unsymmetric if their temporal means are zero or
dierent from zero respectively.
For constant amplitude loading, the response function presents two parts: a transient and a
steady state response (see Fig. 6.8). At the beginning of the loading, in the transient response,
Figure 6.8: Parts of the response for a constant-amplitude harmonic test
Housner "frequency" is still predominant. This "frequency" is close to that corresponding to the
rst rocking angle caused by ground motion.
In the stationary response, the block will move with the frequency of the load. However, it has
long been observed [14],[39],[40],[43], that the transient response is dominant for the stability
of a rocking block. As in previous experimental works [39],[40], the predominant response
observed in the stationary regime was the harmonic (1,1) mode.
6.2. Types of Response and Their Spectra 81
Figure 6.9 shows the stationary part of the responses obtained from constant amplitude tests.
As in [40], the most common type of response is the orbit (1,1), shown in Figure 6.9 a). For
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
6 6.5 7 7.5 8 8.5 9 9.5 10

(
o
)
(s)
a) Harmonic (1,1) Mode
-3
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
15 15.5 16 16.5 17 17.5 18 18.5 19 19.5 20

(
o
)
(s)
b) Subharmonic (2,2) Mode
Figure 6.9: Stationary part of periodic modes. Experimental response.
specimen 2 an orbit was found which resembles the mode (2,2). For this case, an amplitude
of 20mm and a frequency of 2.5Hz were used. Figure 6.9 b) shows the stationary part of the
experimental response of that mode.
For an harmonic loading with constant amplitude, the spectra is shown in Figure 6.10.
0
0.02
0.04
0.06
0.08
0.1
0.12
1 1.5 2 2.5 3 3.5 4
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
H
f
ext
a)
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.18
1.5 2 2.5 3 3.5 4 4.5
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
H
f
ext
b)
Figure 6.10: Fourier spectrum of harmonic loading with constant amplitude; a) specimen 2
under sine with 8mm and 3.3Hz, b) specimen 4 under sine with 5mm and 4.0Hz.
6.2. Types of Response and Their Spectra 82
It can be noticed that, as expected, the main peak corresponds to the external load frequency.
On the other hand, the Housner frequency is still present due to the onset of RM by the rst
pulse. The spectra for the rest of the specimens is similar.
Comparison between theory and experiment
Figure 6.11 shows the comparison of transient and stationary parts of the response of specimen
1 under a sinusoidal load with a constant amplitude of 4mm and a frequency of 3.3Hz. The
tted values of Table E.1 were used. As noticed from the graph, good correlation is achieved
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
9 10 11 12 13 14

(
o
)
(s)
Theory
Experiment
a)
-1
-0.5
0
0.5
1
24 24.5 25 25.5 26 26.5 27

(
o
)
(s)
Theory
Experiment
b)
Figure 6.11: Specimen 1. Comparison theory-experiment for forced regime with constant am-
plitude harmonic load with tted parameters of Table E.1 with 4mm and 3.3Hz a) transient
response, b) stationary response
between the experimental and numerical model. Nevertheless, a phase lag occurs in the transient
response after approximately 12 periods.
On the other hand, qualitatively good agreement is reached for the stationary response which,
despite this phase lag, follows the load history. Similar good agreement is obtained for the rest
of the specimens.
6.2. Types of Response and Their Spectra 83
Harmonic loading. Moduled amplitude
The eects on the response by an harmonic load with constant frequency and a moduled ampli-
tude were studied by means of a Hanning window:
w(n) = 0.5
_
1 cos
_
2n
N 1
__
(6.5)
where N represents the width, in samples and n is an integer with values 0 n N 1.
As stressed above, there is not a well dened Housner frequency for forced motion. However,
it can be still dened the equivalent f
H
as the frequency corresponding to the rst maximum
angle of the response from which rocking is activated. For this case, the spectrum changes sig-
nicantly in two aspects; rst, the secondary frequency -which lies close to the equivalent f
H
-
is now comparable to the main frequency and, second, more peaks besides the main external
frequency appear.
Figure 6.12 depicts the spectrum along with f
H
and the maximum frequencies of specimen 2
for an amplitude of 6mm and two values of the external frequency. Results for the rest of spec-
imens are analogous. The eect of the "windowed" signal on the block can be understood as
0
0.005
0.01
0.015
0.02
0.025
1 1.5 2 2.5 3 3.5 4
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
H
f
ext
a)
0
0.001
0.002
0.003
0.004
0.005
0.006
0.007
0.008
0 0.5 1 1.5 2 2.5 3 3.5 4
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
H
f
max
f
ext
b)
Figure 6.12: Specimen 2. Spectrum of the block response for a hanning sine load with an
amplitude of 6mm and two dierent frequencies: a) 3.3Hz and b) 2.5Hz
that corresponding to an impulse with a given frequency content. As the frequency decreases,
the response resembles the obtained from a single-pulse-type motion.
6.2. Types of Response and Their Spectra 84
On the other hand, for large frequencies, the periodic content of the function becomes predom-
inant and the eect is similar to that observed for constant amplitude tests.
Several cases were analyzed through computer simulation yielding equivalent results. In the
case shown, the frequency is decreased from 3.3Hz to 2.5Hz. While Fig. 6.12 a) still exhibits
a clear constant-amplitude pattern, many new frequencies spring up in Fig. 6.12 b) through a
cascade shape similar to that observed in free RM tests.
Comparison between theory and experiment
The comparison between theory and experiment for a hanning sine test on specimen 1 is shown
in Fig 6.13. The tted values of Table E.1 were used. From this plot, it is clear that reasonable
agreement between theory and experiment is achieved when parameters are allowed to vary.
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
3 4 5 6 7 8 9 10

)
(s)
Theory
Experiment
Figure 6.13: Specimen 1. Comparison theory-experiment for hanning sine with 10mm and
3.3Hz and tted parameters of Table E.1
6.2. Types of Response and Their Spectra 85
Earthquake loading
Generation of input sample functions
The time-series were generated through an automatic procedure, [39], [41], [63]. The procedure
consists on the three following steps:
1. First, a set of samples of stationary Gaussian white noise was generated.
2. Then, those series were multiplied by a variance intensity function of time in order to add
the typical intensity function of accelerograms.
3. Finally, the result was passed through a second order linear lter to impart the desired fre-
quency content. This spectrum was made compatible with the requirements of Eurocode-
8 soil-type A [4],[18].
Figure 6.14 a) shows the time-history of a typical input signal for random loading used in the
present work. The series were normalized in intensity to a maximum PGA value of 2.7m/s
2
and a duration of 14s. Figure 6.14 b) presents the Fourier spectra of 10 input samples in a
logarithmic scale. It can be noticed that the samples are quite regular in the frequency domain.
The main frequencies lie in the range between 0.3Hz and 30Hz with similar weight.
-3
-2
-1
0
1
2
3
0 2 4 6 8 10 12 14
a
(
m
/
s
2
)
(s)
a) Time history
1e-11
1e-10
1e-09
1e-08
1e-07
1e-06
1e-05
1e-04
0.001
0.01
0.1
0.01 0.1 1 10 100
A
m
p
l
i
t
u
d
e
Freq.(Hz)
b) Fourier spectrum
Figure 6.14: Typical input series for a random load test.
6.2. Types of Response and Their Spectra 86
According to this procedure, 500 earthquake records were generated and each sample was con-
sidered equivalent from the probabilistic point of view. The averaged quantities (such as col-
lapse rates) are directly related to the input sample process through Eq. 8.1. Therefore, the basic
properties of a
g
() are analyzed in next.
Stochastic properties of the generating process
The sample functions can be understood as outcomes of a stochastic process with prescribed
characteristics. As illustrated in Figure 6.15, at two sampling times t
1
and t
2
, the statistical
properties of the sample are evaluated.
1
2
3
4
5
6
0 2 4 6 8 10 12 14
t
1
*
t
2
*
7
Figure 6.15: Ensemble of sample input functions
6.2. Types of Response and Their Spectra 87
The number of samples in the ensemble was chosen in order to guarantee the stochastic prop-
erties. Figure 6.16 plots the value of the ensemble mean of the input signal < a
g
> () as
a function of time for two sets of 100 and 275 samples. The maximum deviation found was
0.011. This value allows the reduction of the sample space from 500 to 100 input series.
-0.02
-0.015
-0.01
-0.005
0
0.005
0.01
0.015
0.02
0.025
0 2 4 6 8 10 12 14
<
a
g
>
Time(s)
100 samples
275 samples
Figure 6.16: Ensemble mean of a
g
= a/g as a function of time for two ensembles of 100 and
275 samples. Sampling time = 0.1s
Arelevant issue for the analytical models concerns the stationarity of the input signal. Figure 6.17
presents the histograms for 100 input sample series at dierent sampling times as shown in
Fig. 6.15. A Gaussian t through the least square method has also been plotted. It can be
noticed that the process is fairly Gaussian with independence of the time.
Stationarity of ensemble mean and variance
The sample series were generated with zero mean. However, if quantities are intended to be
derived at a regular sampling frequency, the t of the sampling may alter the result. Figure 6.18
a) presents the ensemble means < a
g
> at sampling intervals of 0.1s and 1s. The maximum
divergence found was 0.015. With this value, although the process is not strictly stationary in
the mean, it can still be considered as such for engineering purposes. The variance of the input
samples < (a
g
)
2
> is investigated for dierent sampling times with intervals of 0.1s and 0.4s.
6.2. Types of Response and Their Spectra 88
0
0.1
0.2
0.3
0.4
0.5
0.6
-3 -2 -1 0 1 2 3
Num. bins: 60, Bin-width: .07695516
GAUSSIAN FIT: =-0.03, =0.99
3.5s
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
-3 -2 -1 0 1 2 3
Num. bins: 60, Bin-width: .08588283
GAUSSIAN FIT: =0.04, =0.88
7.0s
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
-3 -2 -1 0 1 2 3
Num. bins: 60, Bin-width: .08108199
GAUSSIAN FIT: =-0.13, =1.03
10.5s
0
0.2
0.4
0.6
0.8
1
-3 -2 -1 0 1 2 3
Num. bins: 60, Bin-width: .08151400
GAUSSIAN FIT: =-0.07, =0.68
13.9s
Figure 6.17: Histograms of the ensemble at dierent sampling times. Gaussian t is also indi-
cated
From Fig. 6.18 b) it is clear that Var(a
g
) is almost constant in the range 3s-13s with a value of
0.01.
Correlation of input samples
The process is uncorrelated as it can be veried from the recurrence plot (see Section 7.13)
shown in Fig. 6.19 a) and from the auto-correlation function depicted in Fig. 6.19 b). The
recurrence plot of Fig. 6.19 a) was generated with no embedding, a delay of 1 and a threshold
corridor of = 10
2
. From the guidelines exposed in section 7.13 it can be noticed that the input
samples are highly uncorrelated at medium and large time scales. The lack of cluster structures
and white blocks is associated to a high degree of uncorrelation in the input series.
6.2. Types of Response and Their Spectra 89
-0.02
-0.015
-0.01
-0.005
0
0.005
0.01
0.015
0 2 4 6 8 10 12 14
<
a
g
>
Time(s)
Sampling time = 0.1s
Sampling time = 1s
a) Mean
0
0.002
0.004
0.006
0.008
0.01
0.012
0.014
0 2 4 6 8 10 12 14
Time(s)
Sampling time = 0.1s
Sampling time 0.4s
b) Variance
Figure 6.18: Ensemble mean and variance of a
g
= a/g as a function of time. Comparison for
two sampling times
a) Recurrence plot
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
R
(

)
(s)
b) Autocorrelation function
Figure 6.19: Recurrence plot and autocorrelation graph of input series (see section 7.13)
On the other hand, from Fig. 6.19 b) it is clear that the autocorrelation function tends to zero
quite fast. The auto-correlation achieves a value of 10
3
for a separation time of 0.1s. This fact
ensures that the stochastic process can be considered as stationary if sampling rate is greater
than 0.1s.
6.2. Types of Response and Their Spectra 90
Power spectral density
The frequency content of the sample series is implemented through the power spectral density
function S
w
. The function S
w
consists on two main parts:
S
f
=

0, f < f
1
S
1
+ m( f f
1
), f f < f
c
a/( f + b)
2
, f
c
f f
2
0, f > f
2
(6.6)
where m and b are constants and f
1
, f
2
, f
c
are indicated in the Fig. 6.20. The maximum value
lies at the central frequency f
c
= 3.0Hz. The schematic form of S
w
is shown in Fig.6.20 a).
0 5 10 15 20
S
(
f
)
Freq.(Hz)
f
c
f
1
a/(f+b)
2
s1+m(f-f
1
)
f
2
s
max
s
1
s
2
a) Schematic form of S
w
0 5 10 15 20
S
(
w
)
Freq.(Hz)
Design Power Spectral Density
b) Computation of S
w
for 10 input series
Figure 6.20: Power spectral density of the sample input series
Additionally, several power spectral density functions corresponding to dierent sample series
are shown in Fig. 6.20 b) along with the design S
w
.
From the results presented above the main conclusions can be derived for the input loading
characteristics:
The process is ergodic in the mean: < a
g
(t) >

a
g
(t) 0.
The process can be considered Gaussian in intensities.
A constant value of 0.01 can be assumed for the variance.
6.2. Types of Response and Their Spectra 91
Autocorrelation function decreases fast. Therefore, the process becomes highly uncorre-
lated after a short time (typically 0.1s).
Power spectral density function can be obtained from a second order linear lter with the
shape given by Eq. 6.6.
Spectrum for earthquake loading
Figure 6.21 shows the spectrum of the response of Specimens 1 and 3 obtained from a random
load. Graphs for random white noise and for a synthetic earthquake record are reported. As
expected, a bunch of new frequencies appears. The Housner frequency is still reported for
0
0.0005
0.001
0.0015
0.002
0.0025
0.003
0.0035
0.004
0 1 2 3 4 5 6
A
m
p
l
i
t
u
d
e
Freq.(Hz)
f
Hmax
f
max
a)
0
0.002
0.004
0.006
0.008
0.01
0.012
0 0.5 1 1.5 2 2.5 3 3.5
A
m
p
l
i
t
u
d
e
Freq.(Hz)
b)
Figure 6.21: Spectrum of the response under random loading: a) Specimen1. White noise load
with noise-strength=0.5%, b) Specimen 3. Synthetic earthquake record with PGA=0.137g
the white noise case. However, no correlation was detected between f
H
and the predominant
peaks of the response along an intensive computation for several earthquakes and noise-strength
levels.
Comparison between theory and experiment
Figures 6.22 and 6.23 depict the comparison between analytical model and measured response
of specimen 3 under two dierent synthetic earthquakes with a peak ground acceleration (PGA)
6.3. Reproducibility of the Response 92
-4
-3
-2
-1
0
1
2
3
4
5 10 15 20 25

)
(s)
Theory
Experiment
Figure 6.22: Specimen 3. Comparison
theory-experiment for earthquake 2 with
PGA=2.15m/s
2
.
-1.5
-1
-0.5
0
0.5
1
1.5
10 12 14 16 18 20 22 24

)
(s)
Theory
Experiment
Figure 6.23: Specimen 3. Comparison
theory-experiment for earthquake 1 with
PGA=1.72m/s
2
.
of 2.15m/s
2
and 1.72m/s
2
respectively. It is noticed that the theory is able to reproduce quali-
tatively the experimental results. However, as it will be highlighted in the following sections, a
tiny variation of the PGA value can lead to a very dierent response.
6.3 Reproducibility of the Response
Several experiments were conducted under similar conditions in order to study the repeatability
of the rocking phenomenon. Initially, the set consisted on six tests on specimen 1 for a Hanning
sine of 3.3Hz and 7mm. Two of them are depicted in Figure 6.24. Choosing the rst test as
a reference, the dierences (in percentage) of the values of the maximum angle are under 3%
(except for test number 5, which presents an error of 10%).
It has also been noticed that some pair of equal tests appear shifted by an angle of . This fact
is ascribed to the unpredictable nature of the start-of-motion mechanism. The results derived
from this survey led to the conclusion that the repeatability of the response for harmonic tests is
achieved up to slight variations. However, as it will be shown in the following, this observation
does not hold for random motion.
6.3. Reproducibility of the Response 93
Figure 6.24: Experimental responses for two
equal tests. Hanning sine of 7mm and 3.3Hz
Figure 6.25: Specimen 2. Two equal tests for
earthquake 20 and PGA=3.35m/s
2
.
As reported in [39], [43], there is an intrinsic diculty in reproducing random motion tests.
The responses obtained from two equal tests on specimen with earthquake 20 and a PGA equal
to 3.35m/s
2
are depicted in Fig. 6.25. In one of the tests, the specimen resisted the earthquake
load, whereas in the other test it overturned. From Fig. 6.25 it is also noticed that during the
rst six seconds both responses are similar. However, after that time, the two responses became
qualitatively dierent. This behavior is typical of systems whose dynamics are chaotic under
certain conditions.
-4
-2
0
2
4
6 7 8 9 10 11 12 13

(
o
)
(s)
Test 1
Test 2
Test 3
Figure 6.26: Specimen 3. Three equal tests for earthquake 01 and PGA=3.02m/s
2
.
6.4. Classication of Orbits 94
Additional repeatability tests were carried out on the other specimens. For instance, three dif-
ferent experiments with earthquake 1 were conducted on specimen 3 with a maximum PGA of
3.02m/s
2
(Fig. 6.26). In this case, the specimen overturned for the second test.
6.4 Classication of Orbits
Many types responses which are well known in literature ([12],[64]), could not be found by
experimental investigation. Nevertheless, all the types of response previously found by other
authors (for instance, Hogan, [12] or Yim & Lin [15]) can be reproduced by using system of
equations 5.50. These responses can be classied in four main types:
1. Bounded motion.
(a) Periodic.
(b) Quasi-Periodic.
(c) Chaotic.
2. Unbounded motion.
Only periodic and overturning responses where observed during the experimental research. The
reason for this is that, quasi-periodic modes only occur for the unphysical value of = 1.0, i.e,
undamped systems.
On the other hand, several periodic orbits where not found. As in the experimental work carried
out by Tso & Wong [64], these results failed to nd some periodic solutions because they are
unstable.
Periodic response
The rst type of bounded response is the periodic motion. Figures 6.27 and 6.28 show the two
most probable types of orbits. These types are labeled as symmetric (1,1) and (1,3) symmetric-
subharmonic modes respectively. The orbits in the phase plane (where the x-axis and the y-axis
correspond to position and velocity respectively) are depicted in Figs. 6.27 b) and 6.28 b).
6.4. Classication of Orbits 95
a) Time series orbit b) Phase-space trajectory
Figure 6.27: Time series orbit and phase portrait of (1,1) Mode: = 0.245, = 0.925, = 2.42,
= 4, = 1.8 and init.. conds (x
1
(0) = 0.069,x
2
(0) = 0.8,x
3
(0) = 0.0)
a) Time series orbit
b) Phase-space trajectory
Figure 6.28: Time series orbit and phase portrait of (1,3) Mode: = 0.245, = 0.925, = 4,
= 10, = 0.0 and init. conds (x
1
(0) = 0.061,x
2
(0) = 0.061,x
3
(0) = 0.0)
A number of more complicated modes were obtained through suitable variation of initial con-
ditions and parameters. For instance, the symmetric mode (1,7) is shown in Fig. 6.29.
6.4. Classication of Orbits 96
a) Time series orbit b) Phase-space trajectory
Figure 6.29: Time series orbit and phase portrait of (1,7) mode: = 0.245, = 0.925, = 4,
= 14, = 0.0 and init. conds (x
1
(0) = 0.219,x
2
(0) = 0.092,x
3
(0) = 0.0)
Quasi-periodic response
The next type of response, coined as quasi-periodic was rst reported by Yim & Lin [15]. As
the periodic response, quasi-periodic motion is bounded. It is characteristic of systems with two
or more incommensurate frequencies. Typical orbits are shown in Fig. 6.30.
a) Time series orbit b) Phase-space trajectory
Figure 6.30: Time series orbit and phase portrait of quasi-periodic mode: = 0.245, = 1.0,
= 4, = 14 and init. conds (x
1
(0) = 0.075,x
2
(0) = 0.144,x
3
(0) = 0.0)
6.4. Classication of Orbits 97
Chaotic response
Another type of bounded rocking response (also discovered by Yim & Lin [15]) is chaotic mo-
tion. It is characterized by a random-like, unpredictable aspect. The unpredictability arises from
its extreme sensitive dependence on initial conditions. Typical plots of this type are shown in
Figs. 6.31.
a) Time series orbit b) Phase-space trajectory
Figure 6.31: Time series orbit and phase portrait of chaotic response for = 0.245, = 0.925,
l = 1E 5, = 4.0, = 6.1, = 0.0 and init. conds (x
1
(0) = 0.1,x
2
(0) = 0.4,x
3
(0) = 0.0)
Overturning response
Unbounded (overturning) response occurs when the normalized absolute value of the angle;

1
=
||

, reaches 1 and continues to diverge; the response becomes unbounded. The typical
orbits and the phase portraits for this type are shown in Fig. 6.32.
6.5. Summary 98
a) Time series orbit b) Phase-space trajectory
Figure 6.32: Time series orbit and phase portrait of overturning Response for = 0.925, =
1E 5, = 2.42, = 4, = 0.0 and init. conds (x
1
(0) = 0.4,x
2
(0) = 0.0,x
3
(0) = 0.0)
6.5 Summary
The following conclusions can be made from this chapter:
The most common mode observed during the experimental tests under harmonic forcing
with constant amplitude was the periodic (1,1) Mode. Other modes were only found
through numerical simulation.
No quasi-periodic behavior has been found in the experimental tests.
Experimentally, chaotic orbits are not distinguishable from overturning responses.
Good agreement between theory and experiment is only found when the parameter values
are allowed to vary by inverse tting. Then, using constant parameters, the agreement is
very good for free RM and harmonic forcing. Reasonable agreement has also been found
for random loading.
Reproducibility of the response is usually impossible for random loading regime. The
results reported here strength the hypothesis that RM under random load should be un-
derstood as an intrinsically probabilistic phenomenon.
Chapter 7
Stability Analysis
In the last decades data analysis using linear methods were further improved and enriched with
new methods that were derived from chaos theory. The presents contribution applies some
of these new techniques to the RM problem in order to establish an adequate basis for the
understanding and quantication of the rocking motion stability. Some of the results derived
herein will be applied in chapter 8 to the problem of the probability of collapse of single block
structures under random loading. This chapter consists on three steps. First, the statistics of
maximum angles, velocities and energies is investigated using experimental results and numer-
ical simulation. Second, the number of impacts per time is found to be a key parameter for the
stability; its distribution and dependence on system parameters is examined. Third, the mean
exit time, dened as the time for which equals the static collapse angle is investigated nu-
merically.
It was observed that the behavior of the system changes abruptly under certain conditions when
one parameter is slightly modied. Bifurcation thresholds are then found numerically for har-
monic loading under both amplitude and frequency ranges.
A Poincar surface of section (PSS) analysis is performed through an extensive numerical study
along the range of the system parameters. Dierent types of loading with their PSS are dis-
cussed. Finally, recurrence plots and recurrence quantication analysis are used in order to
highlight patterns and inherent structures in the dynamics which are not easy to detect by other
techniques.
99
7.1. Maximum Angles and Overturn 100
7.1 Maximum Angles and Overturn
An issue that is directly related to collapse is the maximum angle reached by the block. In the
present study, the value of |
max
|, that is, the absolute value of the maximum rocking angle, will
be examined under dierent conditions.
Harmonic forcing
The maximum experimental rocking angles for harmonic forcing are reported in Table 7.1 and
Fig. 7.1. In Fig. 7.1 a), the measured
max
are depicted against the external load amplitude for
Constant Sine Amplitude (mm)
(Hz) 03 04 05 06 08 10 12
0.5 NR NR
1.0 NR
1.5 6.08
2.0 NR NR NR 3.04 3.93
2.5 NR NR 2.27 2.33 2.63
3.0 1.36 1.49 1.85 1.87 2.15
3.3 1.35 1.49 1.57 1.42 2.03
5.0 0.95 1.14 1.23
Hanning Sine (Hz)
2.5 NR 2.66
3.0 1.76 1.66 1.81
3.3 1.27 1.53
NR = No rocking
Table 7.1: Specimen 2, absolute value of maximum rocking angle (
o
) for harmonic motion.
dierent frequencies. As expected, it is observed a clear trend;
max
increases with the exter-
nal amplitude. On the other hand, Fig. 7.1 b) shows
max
vs external frequency for dierent
7.1. Maximum Angles and Overturn 101
0
0.5
1
1.5
2
2.5
3
3.5
4
3 4 5 6 7 8 9 10
m
a
x

|

(
o
)
|
TABLE-AMPL (mm)
2.0Hz
2.5Hz
3.0Hz
3.3Hz
a) Harmonic load vs external amplitude
0
0.5
1
1.5
2
2.5
3
3.5
0 1 2 3 4 5
m
a
x

|

(
o
)
|
f(Hz)
3mm
4mm
5mm
6mm
8mm
b) Harmonic load vs external frequency
Figure 7.1: Specimen 2. Experimental maximum angle for harmonic load vs external amplitude
and frequency
values of the load amplitude. For this case, no evident patterns are found, although an overall
conclusion can be derived, namely, resonance peaks appear for certain values of the external
frequency. When such value of frequency is reached, the measured
max
increases abruptly.
Unfortunately, the experimental results are not able to provide sucient data to derive the res-
onance patterns.
Figure 7.2 shows the result of the numerical integration of system of Eqs.5.50 for a grid of
external amplitudes and frequencies. The parameters used were the tted values of specimen 2
and a cosine-type function was chosen as harmonic load.
It can be noticed that for a given external amplitude, maxima occur at dierent values of fre-
quency, which clearly resembles the phenomenom of "resonance" mentioned above.
Several peaks lie in the interval 3.5 4Hz which corresponds to the value of the frequency
parameter p. However, the presence of other peaks at other values of frequency avoids taking
this value as a general rule.
7.1. Maximum Angles and Overturn 102
-2
0
2
4
6
8
10
12
0 1 2 3 4 5 6 7
m
a
x

|

|

(
o
)
Freq. (Hz)
1 mm
2 mm
3 mm
4 mm
5 mm
Figure 7.2: Specimen 2. Maximum angle peaks for constant external amplitude. Numerical
integration of a cosine-type load with an amplitude of 2mm
Random motion
For each earthquake sample, several tests were conducted on specimens 3 and 4 with an in-
creasing value of the PGA. Figure 7.3 plots |
max
| for dierent levels of PGA. The solid line
intersects the average values of |
max
| computed for all the earthquakes with the same PGA. It is
again clear an increase of |
max
| with an increase of loading (PGA).
Distribution of ||
At the same PGA level, tests were conducted on specimens for dierent earthquake records.
From these records, the mean < || > is found at dierent sampling times. Figure 7.4 presents
the experimental values of < || > found for specimen 3 under random loading. From this gure
it can be noticed that there exist a time interval t
2
t
1
for which the value of < || > remains
almost stationary (see Fig. 7.5). After a time equal to t
2
, the angle starts to decrease. The length
of this interval and its origin is severely aected by the external PGA.
7.1. Maximum Angles and Overturn 103
0
1
2
3
4
5
6
7
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6
<
|

m
a
x
(
o
)
|
>
PGA (m/s
2
)
Specimen 3
0
5
10
15
20
3 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 4 4.1 4.2 4.3
<
|

m
a
x
(
o
)
|
>
PGA (m/s
2
)
Specimen 4
Figure 7.3: Experimental maximum angles (absolute value) under random load for Specimens
3 and 4
0
0.2
0.4
0.6
0.8
1
1.2
0 5 10 15 20 25 30
<
|

|
o
>
Time (s)
PGA=1.29m/s
2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 5 10 15 20 25 30 35 40
<
|

|
o
>
Time (s)
PGA=1.72m/s
2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 5 10 15 20 25 30 35 40
<
|

|
o
>
Time (s)
PGA=2.15m/s
2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
0 5 10 15 20 25 30 35 40
<
|

|
o
>
Time (s)
PGA=2.58m/s
2
Figure 7.4: Experimental distribution of ensemble mean of || for specimen 3 under random
loading. Eect of PGA
7.2. Ensemble Average Values of Velocity and Energy 104
0
0.5
1
1.5
2
2.5
3
0 5 10 15 20 25 30 35 40
<
|

|
o
>
Time (s)
STATIONARY INTERVAL
t1 t2
Figure 7.5: Schematic graph of the experimental stationary limit of < || > under random load-
ing
The stationary limit of < || > can be computed as follows. First, the histogram of values
of < || > is built. Then, the value at which this histogram has a maximum is computed. This
value, corresponding to the most probable value of < || >, is regarded as the stationary limit.
Once the stationary value of < || > is given, the coordinates t
1
and t
2
are found by inverse root
nding. Figure 7.6 plots the values of this stationary range for dierent PGA values. It can be
noticed that the value of < || > increases continously with PGA.
7.2 Ensemble Average Values of Velocity and Energy
The velocity and energy of the rocking block are random variables whose distribution and max-
ima are key quantities for the correct understanding of RM collapse. This section addresses
the ensemble distribution of the velocity and energy along with their dependence on PGA and
system parameters.
Distribution of velocities
As it was found for the angles, it has been observed that < v > lies within an interval where its
value remains quasi stationary. Figure 7.7 plots the ensemble means of the experimental veloc-
7.2. Ensemble Average Values of Velocity and Energy 105
30
35
40
45
50
55
60
65
70
75
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2
s
t
a
t

l
i
m
i
t

o
f

<

>

(
o
)
PGA (m/s
2
)
Figure 7.6: Experimental stationary limit of < || > for specimen 3 under random loading found
as the most probable value of < || >. West acceleration (see Eq. 6.2) is indicated by a straight
line.
ity measured with specimen 3 under two dierent levels of PGA for dierent input samples.
0
1
2
3
4
5
6
7
8
0 5 10 15 20 25 30 35
<
|d

/
d

|>

(
(o
) /
s
)
Time(s)
PGA = 2.15m/s
2
0
1
2
3
4
5
6
7
8
9
10
0 5 10 15 20 25 30 35
<
|d

/
d

|>


(
(o
) /
s
)
Time(s)
PGA = 2.58m/s
2
Figure 7.7: Specimen 3. Experimental distribution of velocity
It is noticed that the interval for which experimental velocities remain stationary is signicantly
larger than the corresponding interval to angles (Fig. 7.4).
On the other hand, Fig. 7.8 depicts the stationary limit of the ensemble average of < v > for
the experimental results of specimen 3 as function of the external PGA. As expected, < v >
increases continuously with PGA.
7.2. Ensemble Average Values of Velocity and Energy 106
It must be highlighted, however, that, as it is shown in Figs. 7.6 and 7.8, there is a motion
at lower values than the West acceleration. This fact is due to the dispersion (standard devia-
tion) in the a
w
value, which was obtained from Eq. 6.2 with the tted parameters of Table E.1.
5.6
5.8
6
6.2
6.4
6.6
6.8
7
7.2
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2
<
v
>

(
r
a
d
/
s
)
PGA (m/s
2
)
Figure 7.8: Dependence of experimental < v > of specimen 3 with PGA. West acceleration is
indicated with a straight line
Figure 7.8 shows the experimental values of the stationary limit of < v > for specimen 3. As it
was expected, the stationary limit of the velocity increased with PGA.
Distribution of energies
In order to characterize the energies of the orbits, rst the experimental results for earthquake
loading are examined. Figure 7.9 shows the nondimensional potential, kinetic and mechanical
energies measured from specimen 3 under a random loading test.
From this gure it can be noticed that the mechanical energy remains almost constant during
impacts, and it changes abruptly (in a Dirac-delta type fashion) at the impact times.
Since time intervals for which the values of || and v remain almost stationary were found, there
must also exist such stationary intervals for the mechanical energy. Figure 7.10 depicts the me-
7.2. Ensemble Average Values of Velocity and Energy 107
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
13 13.1 13.2 13.3 13.4 13.5 13.6 13.7 13.8
Time(s)
||(
o
)
K (x 200)
U (x 200)
E (x 200)
Figure 7.9: Experimental non-dimensional potential (U), kinetic (K) and mechanical (E) ener-
gies of specimen 3 under random loading
chanical energy of specimen 3 for random loading tests with increasing PGA values. From this
graph it can be noticed that, for a given earthquake record higher PGA values do not necessarily
imply larger energies. However, it will be shown at the end of this section through numerical
simulation that, in the stationary limit of the ensemble, the mean of the mechanical energy in-
creases with PGA.
The eect of the slenderness ratio in the ensemble mean of the nondimensional energy e =
E/I p
2
is monitored by numerically integrating 100 earthquake records for a xed level of PGA
and dierent . The remaining parameters are xed to the values: p = 3.6s
1
, = 0.925.
Figure 7.11 presents the results from that analysis corresponding to two = h/b values with a
PGA of 4.05m/s
2
. It can be noticed that < e > increases with the slenderness ratio. This means,
that slender blocks accumulate more energy and, hence, are more likely to collapse.
7.2. Ensemble Average Values of Velocity and Energy 108
0
0.002
0.004
0.006
0.008
0.01
2 4 6 8 10 12 14 16 18 20
E
n
e
r
g
y
/
I
p
2
Time (s)
PGA= 0.81m/s
2
PGA=1.08 m/s
2
PGA=1.35 m/s
2
PGA= 1.62 m/s
2
PGA= 1.89 m/s
2
PGA= 2.16 m/s
2
Figure 7.10: Experimental mechanical energy of specimen 3 under random loading for dierent
PGA values.
0
0.001
0.002
0.003
0.004
0.005
0.006
0.007
0.008
0.009
0 200 400 600 800 1000 1200
<
e
>
Time (s)
=3
=4
Figure 7.11: Ensemble average mechanical energy for two values
7.2. Ensemble Average Values of Velocity and Energy 109
The frequency parameter is monitored in a similar way; For each p value, the angle is inte-
grated for all the earthquake records with xed PGA, and . Figure 7.12 shows the results
corresponding to the values: p = 2.0s
1
, p = 3.0s
1
, p = 4.0s
1
and p = 5.0s
1
under a PGA
of 2.7m/s
2
, = 4 and = 0.925. Although strong nonlinear eects are present, on average,
0
0.0005
0.001
0.0015
0.002
0.0025
0 2 4 6 8 10 12 14
E
n
e
r
g
y

/

I
p
2
Time (s)
p=2.0
p=3.0
p=4.0
p=5.0
Figure 7.12: Stationary limit of < e > vs p
the mechanical energy decreases with p (and, hence, it increases with R). For a given PGA and
aspect ratio () values, two blocks with dierent size (R) carry a dierent amount of mechanical
energy.
Finally, the eect of the damping parameter on the stationary limit of the ensemble mean of
mechanical energy is examined. For this case, the PGA used was 4.05m/s
2
whereas the remain-
ing parameters were = 4, p = 3.6s
1
. It has been found that < e > decreases exponentially
with for a given value of the external PGA. Figure 7.13 depicts the stationary limits of < e >
for dierent values of along with the curve a exp(b) in order to highlight the exponential
behavior of < e > with . The analysis of the stationary limit of the ensemble, labeled as e,
leads to the following conclusions:
7.2. Ensemble Average Values of Velocity and Energy 110
0
0.002
0.004
0.006
0.008
0.01
0.65 0.7 0.75 0.8 0.85 0.9 0.95
S
t
a
t
i
o
n
a
r
y

l
i
m
i
t

o
f

<
e
>

Numerical Simulation
Exponential fit
Figure 7.13: Stationary limit of < e > vs
e increases with PGA once the threshold value of the west acceleration a
w
has been
reached.
For a given PGA value, e increases with , decreases with p and increases exponentially
with .
On the other hand, once the lower bound limit of a
w
is reached, a linear dependence of < e >
and PGA can be found. These observations can be condensed in the following model for the
stationary limit of the ensemble mean mechanical energy as a function of the external PGA:
e = (p, , )(PGA a
w
) (7.1)
where is a constant which is subjected to the conditions exposed above. Fig. 7.14 plots the
stationary limit of < e > as a function of the external PGA. A linear t has also been included
in the graph in order to highlight the linear correlation present.
7.3. Correlation Maps for Experimental Maximum Angles 111
-0.002
0
0.002
0.004
0.006
0.008
0.01
0.012
2.5 3 3.5 4 4.5
S
t
a
t
i
o
n
a
r
y

l
i
m
i
t

o
f

<
e
>
PGA (m/s
2
)
Numerical Simulation
Linear FIT
Figure 7.14: Stationary limit of < e > vs PGA (, p and kept constant).
7.3 Correlation Maps for Experimental Maximum Angles
In order to investigate the dependence between maximum angles and their possible correlations,
the values of the maximum of the absolute values of the angles; m
n+1
max(|
n+1
|) against
m
n
max(|
n
|) are depicted. If N observations of two sets of data, x
i
, i = 1..N, y
i
, i = 1..N are
available, the linear correlation coecient is dened as:
Corr(x, y) =
N

i=1
(x
i
< x >)(y
i
< y >)
N
x

y
(7.2)
where < x >,
x
, and < y >,
y
are the mean and standard deviations of x and y respectively. In
the following computations the square of Corr(m
n
, m
n+1
) is computed for each type of external
action.
Free RM
Figure 7.15 reports the experimental results of the correlation between consecutive maximum
for free RM regime for the four specimens. As it is clear from Fig. 7.15 there exist a tendency
7.3. Correlation Maps for Experimental Maximum Angles 112
0
0.5
1
1.5
2
2.5
3
0 0.5 1 1.5 2 2.5 3
m
n
+
1

(
o
)
m
n
(
o
)
a)
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
m
n
+
1

(
o
)
m
n
(
o
)
b)
1
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6
m
n
+
1

(
o
)
m
n
(
o
)
c)
0
0.5
1
1.5
2
2.5
0 0.5 1 1.5 2 2.5
m
n
+
1

(
o
)
m
n
(
o
)
d)
Figure 7.15: Map of m
n
vs m
n+1
for free rocking motion; a) specimen 1; b) specimen 2; c)
specimen 3; d) specimen 4.
for the points to lie at the right-hand side of the main diagonal. This drift is due to damping and
it can be used to estimate the coecient through numerical simulation. On the other hand,
Table 7.2 reports the linear correlation coecient for free rocking motion found in the tests. As
it is clear, there exist a stronger linear relationship between consecutive maxima. The peculiar
response of specimen 3, which exhibits the lower correlation ratio, should be highlighted.
The results found above are in agreement with Housner theory for free RM, where an analyt-
ical expression for consecutive maxima can be derived due to conservation of energy between
impacts (Eq. 2.17).
7.3. Correlation Maps for Experimental Maximum Angles 113
Specimen Corr
2
(m
n
, m
n+1
)
1 0.994
2 0.981
3 0.914
4 0.982
Table 7.2: Experimental linear correlation coecient for m
n
and m
n+1
. Free rocking
Harmonic forcing. Constant amplitude
The same computation is performed for the forced motion in order to highlight possible patterns
in the dynamics as the parameter values change. In this case, however, the nonlinear relationship
between consecutive maxima is obvious for the dependence between m
n+1
and m
n
.
Eect of external amplitude
In Fig. 7.16 the correlation maps obtained from harmonic tests on specimen 2 are shown for
increasing values of the external amplitude. In general, the correlation rapidly scatters away
from the main diagonal as the number of maxima (and hence, time) increases. It must be
noticed that for an amplitude equal to 10mm, there exist a functional relationship between m
n
and m
n+1
of the type:
(m
n+1
b)
2
+ (m
n
a)
2
= ( f ())
2
(7.3)
where "a" and "b" are the centers of the spiral and f () is an increasing function of time. This
relationship leads to the spiral-shaped patter shown in Fig. 7.16 d).
Table 7.3 summarizes the value of Corr(m
n
, m
n+1
) for a sinusoidal action of constant frequency
equal to 2.0Hz and varying amplitude. It is stated that the linear correlation between maxima
increases with the external amplitude. However, as stated above, the system is clearly nonlinear
and measures from Corr(m
n
, m
n+1
) are to be taken carefully.
7.3. Correlation Maps for Experimental Maximum Angles 114
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
m
n
+
1

(
o
)
m
n
(
o
)
a)
4mm
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
m
n
+
1

(
o
)
m
n
(
o
)
b)
6mm
0
0.5
1
1.5
2
2.5
3
3.5
0 0.5 1 1.5 2 2.5 3 3.5
m
n
+
1

(
o
)
m
n
(
o
)
c)
8mm
0
0.5
1
1.5
2
2.5
3
3.5
4
0 0.5 1 1.5 2 2.5 3 3.5 4
m
n
+
1

(
o
)
m
n
(
o
)
d)
10mm
Figure 7.16: Map of m
n
vs m
n+1
for Harmonic load. Experiment
Amplitude (mm) Corr
2
(m
n
, m
n+1
)
4 0.067
6 0.550
8 0.722
10 0.736
Table 7.3: Experimental linear correlation coecient for maxima under harmonic loading. Ef-
fect of external amplitude
7.3. Correlation Maps for Experimental Maximum Angles 115
The fact that the correlation between maxima increases with the external amplitude is inter-
preted as follows: At very low levels of angle displacement the block impacts many times with
the seismic table and the system dynamics are dominated by noisy behavior. As a consequence,
the correlation is low. On the other hand, if large displacements dominate the system, the system
behaves in a more deterministic fashion (noise is a secondary eect) and correlation grows. As
it will be shown in section 7.13, for random loading, the situation is dierent and the correlation
decreases as the external load values increase.
Eect of external frequency
The eect of external frequency is examined through an equivalent procedure. Figure 7.17
shows the correlation maps for specimen 2 for an external maximum amplitude of 6mm and for
dierent external frequencies. It is observed that, contrary to what it was found for the external
amplitude case, the correlation coecient remains almost constant for the whole range of fre-
quencies. The values scatter away from main diagonal independently of the external frequency.
On the other hand, for the value of 5.0Hz, the points in the (m
n
, m
n+1
) map accumulate in the
maxima equal to 0.62
o
. This fact indicates that for this value of the external frequency, the
solution lies in a stationary regime.
The values of the linear correlation coecient for dierent external frequency values are col-
lected in Table 7.4. The results obtained with moduled amplitude (hanning sine functions) were
Frequency (Hz) Corr
2
(m
n
, m
n+1
)
2.5 0.671
3.0 0.741
3.3 0.743
5.0 0.605
Table 7.4: Experimental linear correlation coecient for maxima under harmonic loading. Ef-
fect of external frequency
7.3. Correlation Maps for Experimental Maximum Angles 116
0
0.5
1
1.5
2
2.5
3
0 0.5 1 1.5 2 2.5 3
m
n
+
1

(
o
)
m
n
(
o
)
a)
2.5Hz
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
m
n
+
1

(
o
)
m
n
(
o
)
b)
3.0Hz
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
m
n
+
1

(
o
)
m
n
(
o
)
c)
3.3Hz
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 0.2 0.4 0.6 0.8 1 1.2 1.4
m
n
+
1

(
o
)
m
n
(
o
)
d)
5.0Hz
Figure 7.17: Map of m
n
vs m
n+1
for Harmonic load. Experiment
equivalent to those corresponding to constant amplitude tests. However, for random loading (as
it is customary), the behavior is qualitatively dierent.
Random loading. Correlation maps of maxima
Finally, the correlation of maxima is investigated for random loading. Figure 7.18 presents the
correlation maps for specimen 3 under the same input record for dierent PGAs ranging from
0.81m/s
2
to 2.16m/s
2
.
7.3. Correlation Maps for Experimental Maximum Angles 117
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
m
n
+
1

(
o
)
m
n
(
o
)
a)
PGA = 0.81m/s
2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 0.2 0.4 0.6 0.8 1 1.2 1.4
m
n
+
1

(
o
)
m
n
(
o
)
b)
PGA = 1.08m/s
2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
m
n
+
1

(
o
)
m
n
(
o
)
c)
PGA = 1.35m/s
2
0
0.5
1
1.5
2
2.5
3
3.5
4
0 0.5 1 1.5 2 2.5 3 3.5 4
m
n
+
1

(
o
)
m
n
(
o
)
d)
PGA = 1.62m/s
2
0
1
2
3
4
5
6
0 1 2 3 4 5 6
m
n
+
1

(
o
)
m
n
(
o
)
e)
PGA = 1.89m/s
2
0
2
4
6
8
10
12
14
16
18
0 2 4 6 8 10 12 14 16 18
m
n
+
1

(
o
)
m
n
(
o
)
f)
PGA = 2.16m/s
2
Figure 7.18: Map of m
n
vs m
n+1
for random load. Experiment tests on specimen 3 with dierent
PGA values
7.3. Correlation Maps for Experimental Maximum Angles 118
There are two states of the system which aect the correlations between maxima:
Micro-RM There are fast oscillations with very small amplitudes about = 0. The systems
dynamics are governed by a noisy-type behavior. There is an intensive exchange of energy
between the system and the environment. The correlation is low.
Transient Free RM The system goes through states which are similar to the free RM motion.
High correlation is achieved.
If the external earthquake is such that transient free RM states appear, the global correlation
increases signicantly.
Table 7.5 summarizes the values for the experimental linear correlation coecient correspond-
ing to Figures 7.18. The high values obtained are due to the presence of this transient free RM
PGA (m/s
2
) Corr
2
(m
n
, m
n+1
)
0.81 0.994
1.08 0.779
1.35 0.794
1.62 0.960
1.89 0.978
2.16 0.994
Table 7.5: Experimental linear correlation coecient for maxima under random loading. Spec-
imen 3 under dierent PGA values
states.
Estimate of maximum correlation length
The results presented so far dealt with the correlation of two consecutive maxima. The next step
is the estimation of the correlation length of maxima, i.e. the number of extrema for which the
correlation function lies within a given -neighborhood.
Figure 7.19 plots several autocorrelation functions for the maxima of obtained from specimen
7.4. Number of Impacts 119
3 under random loading. In this gure, the abscissa represent the correlation length in terms
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 10 20 30 40 50 60 70 80
R
(
L
)
Peak distance-L
PGA = 0.81 m/s
2
PGA = 1.08 m/s
2
PGA = 1.35 m/s
2
PGA = 1.62 m/s
2
PGA = 1.89 m/s
2
PGA = 2.16 m/s
2
Figure 7.19: Autocorrelation function of maxima of ||. Random loading
of the number of consecutive peaks (a distance of 20, for instance, means 20 peaks within
the interval of correlation). It can be noticed that the autocorrelation function of maxima is a
rapidly decreasing function with distance between maxima. Therefore, maxima becomes soon
uncorrelated as the number of impacts between them increases.
7.4 Number of Impacts
The number of impacts, N(), the block undergoes for a given time is a key parameter in order
to understand the stability of the RM dynamics. From Eq. 2.17 it is clear that the number of
impacts is directly related to the energy dissipation rate of the system. Obviously, the number
of impacts is an increasing function of time. In order to eliminate the constant increasing rate
with time, the number of impacts per period is divided by the time.
Figure 7.20 shows the number of impact per time
1
along with the non-dimensional energy of
1
The number of impacts accumulated at a given time is divided by the time
7.4. Number of Impacts 120
the system found in experiment. The times series of () and several impacts times are also
indicated by straight lines. It is clear how the energy of the system is directly related to the
-3
0
3
(
o
)
0
2
4
6
8
Number of impacts / time
0
0 0.5 1 1.5 2 2.5 3 3.5 4
(s)
Energy / Ip
2
0
0 0.5 1 1.5 2 2.5 3 3.5 4
(s)
Energy / Ip
2
Figure 7.20: Experimental number of impacts per time for free rocking regime of specimen 1
impacts. This eect makes the energetic mechanism of the RM behave through a particular
way, which is completely dierent of continuous based models.
Experimental results
First, the experimental number of impacts per time is found for the basic types of loading used.
Figures 7.21 to 7.24 present those results. From this survey it is clear that the function N()
N()/ changes abruptly when rocking is initiated. This triggering time has been marked by a
vertical straight line.
Sine-sweep tests
Figures 7.21 and 7.22 show the experimental values of N

for sine-sweep tests on specimen 1. It


is noticed that the prole of the function N()/ for run-up (Fig. 7.21) and run-down (Fig. 7.22)
tests are completely dierent. This fact evidences the strong dependence of the number of
7.4. Number of Impacts 121
impacts with the external frequency.
-2
-1
0
1
2
0 5 10 15 20 25 30
(s)
(
o
)
8
9
10
11
12
13
14
15
16
17
Number of impacts / time
Figure 7.21: Specimen 1. Experimental number of impacts as function of time. RUN-UP test
for 5mm of maximum table displacement, frequency interval (0.5Hz 5.0Hz)
In the run-up test the external frequency is increased from 0.5Hz to 5.0Hz continuously for
a constant maximum table displacement of 5mm. This load produces that, after a time of 9s
approximately, the block is set into rocking. Since the external frequency is still increasing,
the block is unable to "admit" this higher values of frequency and its motion becomes partially
decoupled from the external loading. This facts leads to a continuous decrease of N()/ from
that time as it is shown in Fig. 7.21.
On the other hand, run-down tests smoothly decreases the external frequency from 5.0Hz to
0.5Hz for a constant table displacement of 5mm. For this case the response of the block is
completely dierent than that corresponding to run-up test. In this situation, as it is shown in
Fig, 6.1, the block augments its energy continuously from the beginning of the motion. Since
the frequency is decreasing, the system undergoes into a regime with high energy and slow
external motion. This state is maintained in time resulting in an smoothly increasing value of
N()/.
7.4. Number of Impacts 122
-4
-3
-2
-1
0
1
2
3
4
5 10 15 20 25 30 35 40 45 50
(s)
(
o
)
0
1
2
3
4
5
6
7
5 10 15 20 25 30 35 40 45 50
Number of impacts / time
Figure 7.22: Specimen 1. Experimental number of impacts as function of time. RUN-DOWN
test for 5mm of maximum table displacement, frequency interval (5.0Hz 0.5Hz)
Harmonic forcing
The experimental results for N

of specimen 1 under harmonic loading are shown in Fig. 7.23.


A maximum table displacement of 4mm and a external driving frequency of 3.3Hz were used.
It is noticed that when the RM is initiated (about 9s), N

decreases smoothly with time.


2
1
-2
0
-1
5 10 15 20 25
(s)
(
o
)
8
14
12
10
Number of impacts / time
Figure 7.23: Specimen 1. Experimental number of impacts as function of time. Harmonic
loading with an amplitude of 4mm and a frequency of 3.3Hz
7.4. Number of Impacts 123
Random loading
Experimental results from random loading tests are also used to compute the number of impacts.
Figure 7.24 presents the experimental results of N

for specimen 4 under random loading with


a PGA value of 1.7m/s
2
. As it was the case for harmonic and sine sweep tests, N

decreases
4
-4
0
0 5 10 15 20
(s)
(
o
)
13
11
9
7
5
Number of impacts / time
Figure 7.24: Specimen 4. Experimental number of impacts as function of time. Random load-
ing: earthquake record 01, PGA=1.72m/s
2
abruptly once RM is initiated. However, for this case it is noticed that after an interval of
duration of approximately 12 seconds, N

commences to increase. This is due to the presence


of free RM behavior once the load is unable to maintain the stationary motion. As it was the
case for free RM regime (Fig. 7.20), N

smoothly increases with time.


Estimation of the number of impacts through numerical simulation
A possible model for the number of impact per time N

N()/ as function of the system


parameters and external PGA is addressed now. Through numerical computation the following
observations can be made:
1. N

decreases as a power of (PGA a


w
).
2. N

increases as a power of and, hence, it decays as a power of .


7.4. Number of Impacts 124
3. N

depends strongly on p in a nonlinear fashion; rst, it increases with p until a maximum


value is reached, then rapidly decreases.
4. N

does not change signicantly with over a the range 0.7 1.0.
The dependences of N

with the parameters is shown in the following.


Eect of PGA
0
10
20
30
40
50
60
70
80
90
0 2 4 6 8 10 12 14
N
u
m
b
e
r

o
f

i
m
p
a
c
t
s

/

t
i
m
e
(s)
PGA=2.7m/s
2
PGA=5.4m/s
2
Figure 7.25: Number of impacts as function of time. Eect of PGA for random loading under
numerical simulation
Figure 7.25 plots the values of N()/ as function of time. For this study, several sets of input
series were used for dierent PGA values. Then, for each PGA, several numerical integrations
were carried out for dierent . As it can noticed from Fig. 7.25, there exist a stationary limit
N

;
n
p
lm

, (7.4)
which is independent of time. It can be seen that the number of impacts per time do not decrease
in the same way for dierent PGA values; larger PGAs lead to a faster decrease. This fact is
reasonable since large PGA values turn the block into a more energetic state at which impacts
occur less frequently.
7.4. Number of Impacts 125
Eect of
The eect of two values of slenderness ratio on the impacts rate N

is shown in Fig. 7.26 a).


On the other hand, the stationary limit n
p
as function of the slenderness ratio is examined and
0
10
20
30
40
50
60
70
80
90
0 2 4 6 8 10 12 14
N
u
m
b
e
r

o
f

i
m
p
a
c
t
s

/

t
i
m
e
(s)
=4
=9
a) N

for two = 4, 9
0
10
20
30
40
50
60
70
80
90
100
1 2 3 4 5 6 7 8 9
n
p

(
s
-
1
)

b) n
p
as function of
Figure 7.26: Stationary limit of the number of impacts per time as function of
shown in Fig.7.26 b). From both graphs it is clear that the number of impacts per time (and its
stationary value) decreases with the slenderness ratio. This result -which has not been reported
in the literature- is considered as much relevant. Slender blocks have lower impact rates.
Eect of p
Figure 7.27 a) shows the solutions of N()/ as function of the frequency parameter p. The
stationary limit n
p
is shown in Fig.7.27 b). The strong nonlinearity of the relationship of the
number of impacts with p is clear. First, n
p
increases smoothly with p until a maximum is
reached. But, as p exceeds this limit value of 5.5 s
1
(approx.) the number of impacts decays
fast. This fact is interpreted as a new manifestation of the so called scale eect, since p is
directly related with the scale (for rectangular blocks p
2
= 3g/4R).
7.4. Number of Impacts 126
0
10
20
30
40
50
60
70
80
90
0 2 4 6 8 10 12 14
N
u
m
b
e
r

o
f

i
m
p
a
c
t
s

/

t
i
m
e
(s)
p=4.0
p=2.0
a) N

for p = 2, 4
0
10
20
30
40
50
60
70
80
90
1 2 3 4 5 6 7 8
n
p

(
s
-
1
)
p (s
-1
)
b) n
p
as function of p
Figure 7.27: Stationary limit of the number of impacts per time as function of p
Eect of
The eect of the damping factor on the impacts rate are shown in Fig. 7.28 a) and b). From
this graph it is noticed that the dependence of N()/ with is not as sensitive as it was found
with the rest of parameters. The values of N

(Fig. 7.28 a)) do not exhibit abrupt changes when


the value of is varied. On the other hand, the stationary limit of n
p
against (Fig. 7.28 b))
0
10
20
30
40
50
60
70
80
90
0 2 4 6 8 10 12 14
N
u
m
b
e
r

o
f

i
m
p
a
c
t
s

/

t
i
m
e
(s)
= 0.8
= 0.9
= 1.0
5
10
15
20
25
30
35
40
45
50
0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
n
p

(
s
-
1
)

Figure 7.28: Stationary limit of the number of impacts per time as function of
decreases almost linearly with . If increases (low damping), the orbital amplitudes are larger
and, within a given interval of time, the impact rate is reduced.
Housner found a analytical relationship between the impact times and (Eq. E.3 of annex E)
7.4. Number of Impacts 127
for free RM regime [7]. In this case, an equivalent numerical relationship has been obtained for
random loading.
Summary: model for n
p
According to the results presented in the preceding analysis of the following model is proposed
for the stationary limit of N():
n
p
= n
p0
+
c
1
p

PGA a
w
(7.5)
where n
p0
and c
1
are two constants which depend on the system parameters (, ).
Figure 7.29 plots the mean values (over an ensemble of identical input samples) of n
p
for dif-
ferent PGA values ranging from the West acceleration to 11m/s
2
. It is noticed that, as soon
0
10
20
30
40
50
60
0 2 4 6 8 10 12
n
p

(
s
-
1
)
PGA (m/s
2
)
NO RM
NUMERICAL SIMULATION
MODEL
Figure 7.29: Stationary limit of the number of impacts per time as function of external PGA
as motion is initiated, n
p
abruptly reaches a value of approximately 17 impacts per second at a
PGA of 4m/s
2
. Then n
p
decays smoothly from this value as the external PGA increases.
7.5. Evaluation of First Passage Time 128
7.5 Evaluation of First Passage Time
From a probabilistic point of view, a relevant parameter related to the probability of collapse is
the time at which the absolute value of the rocking angle exceeds the static collapse angle .
This time is called First Passage Time (FPT), dened as the time

:
| || (

) | (7.6)
where is a tolerance interval.
In order to estimate the FPT and its dependence with PGA, an ensemble of 100 earthquake
records were integrated for dierent PGA values. Then the FPT is computed by averaging over
the ensemble with equal PGA. From this survey the following functional relation was found for
the mean FPT:
< FPT >=
c

PGA a
w
(7.7)
where c is a quantity which may depend on the system parameters; (, p, ). Figure 7.30 depicts
the values obtained from numerical simulation along with a least square t of numerical data to
Eq. 7.7. It must be highlighted that < FPT >, as dened above, is the mean exit time for the
exceedance of the critical collapse angle . As it will be shown in section 8.4, when a collapse
criteria is to be derived, the expression of < FPT > may be inadequate.
7.6 Stability in the Amplitude-Frequency Range
The analysis of the stability under the simplied model of an harmonic force in the amplitude-
frequency, (or ) parameter space has shown to be extremely useful in the understanding of
RM dynamics [14],[15]. In this model, the assumed external load has the form:
a = gcos(
pt
2
+ ) (7.8)
where and are the non-dimensional amplitude and frequency parameters introduced in chap-
ter 5 and is the external phase of the action.
For the present study, will carry the limit values of 0 and /2 in order to investigate the ef-
fect of functions with a cosine and sine shapes. The harmonic model is addressed in this chapter
7.6. Stability in the Amplitude-Frequency Range 129
2
4
6
8
10
12
14
0 5 10 15 20 25 30
<
F
P
T
>

(
s
)
PGA (m/s
2
)
NUMERICAL SIMULATION
FIT: c/(PGA-a
w
)
1/2
Figure 7.30: Ensemble average of First Passage Times for random loading. Numerical simula-
tion against least square tting using Eq. 7.7
through an intensive numerical study of the region.
As suggested in section 7.1 and Fig. 7.1, in RM an eect analogous to the resonance phe-
nomenon exists. The main dierence with standard resonance -where frequency is assumed
independent of external amplitude- is that in this case, if the overturning frequency appears for
a given external amplitude, it will not necessarily stand for other values of the amplitude. In
RM the load amplitude and the response frequency are coupled and, hence, instability must be
investigated in amplitude-frequency pairs.
Power balance
An issue which is directly related to the existence and stability of steady state solutions is the
energetic balance of the system. From Eqs. 5.16 and 5.41 a time-dependent potential A(r, , )
can be built by requiring:
A(r, , ) = U
_
Q
e
r
dr (7.9)
7.6. Stability in the Amplitude-Frequency Range 130
which results into:
A(r, , ) = MgR[cos( r) cos() + a
g
() cos() sin( r)] (7.10)
where "a
g
() = a()/g" is the normalized external acceleration.
The energy is given by:
E = K(r

) + V
b
(r) + A(r, , ) (7.11)
where K = Ir
2
/2 and V
b
(r) = l
2
/(2Ir
2
) is the potential barrier at r = 0. By absorbing the barrier
term into a time-dependent eective potential: V
e
(r, ) = V
b
+ A, the energy is rearranged as:
E(r, r

, ) = K(r

) + V
e
(r, ) (7.12)
The power dissipated by the system is given by [52]:
P =
dE
d
= r

Q
d
r
+
V
e

= r

Q
d
r
+
A

(7.13)
By use of Eqs. 5.35, C.17 and 7.10 it is found, that
e = ln() r
2

j
(t t
j
) + a
g
cos()sin( r) (7.14)
where e E/I p
2
and the sum is extended over the times of impact "t
j
".
Now, this normalized power P = e is averaged over an external period "T
ext
" (the bar symbol
indicates time average):
P = ln() r
2

j
(t t
j
) + a
g
cos()sin( r) (7.15)
The rst term on the right-hand side of the above equation represents the time average rate of
energy loss due to impacts
2
whereas the second term represents the power supplied by the exter-
nal action over a period. A steady-state is achieved when these terms balance. On the contrary,
resonance occurs if the damping term is not able to balance the additional energy supplied by
the external action.
Figure 7.31 shows the instability thresholds of specimen 1 obtained from numerical integra-
tion of the orbits for xed values of the external amplitude and frequency. As the frequency
2
Notice that this term is always negative or zero since 0 1 and, hence, ln() 0.
7.6. Stability in the Amplitude-Frequency Range 131
0
5
10
15
20
0 5 10 15 20 25 30
|

|
m
a
x
(
o
)
PGA (m/s
2
)
O
V
E
R
T
U
R
N
a) Bifurcation in the amplitude range
0
5
10
15
20
0 5 10 15 20 25 30 35
|

|
m
a
x
(
o
)
PGA (m/s
2
)
O
V
E
R
T
U
R
N
a) Bifurcation in the frequency range
Figure 7.31: Specimen 1. Numerical overturn thresholds for external amplitude and frequency
under harmonic forcing regime: a) Constant amplitude of 1.0m/s
2
b) Constant frequency of
f = 0.46Hz
decreases in Fig.7.31 a), the behavior changes abruptly when the external frequency reaches
a value of 4.0Hz. Lower values of frequency lead to overturn. Similar results are reported in
Fig. 7.31 b) for the amplitude case.
In the following, stability maps are addressed by taking into account the eect of the set
of parameters (), p, on the stability of the response. The following grid of values for
was chosen during the computations:
0.0 8.0 with step: 0.1
0.0 20.0 with step: 0.1
For each (, ) the orbit is integrated and the absolute value of maximum angle is computed.
Maps are presented in terms of the dimensional amplitude (g) and frequency (p/2) as
contour plots. A greyscale was chosen, where white represents overturn (|| /2) and, in
the opposite limit, orbits non developing into RM are pixelized with black. Between these
limits, grey levels indicate instability. As it is usual, one parameter varies while the rest remain
7.6. Stability in the Amplitude-Frequency Range 132
constant. The reference values of the parameters have been selected as the most used by other
authors (for instance, [12],[14],[39]), namely, ( = 0.245, p = 3.6, = 0.925, = 0.0).
Eect of the slenderness ratio
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
2 4 6 8 10 12 14
1
2
3
4
5
6
7
= 3
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
2 4 6 8 10 12 14
1
2
3
4
5
6
7
= 6
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
2 4 6 8 10 12 14
1
2
3
4
5
6
7
= 9
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
2 4 6 8 10 12 14
1
2
3
4
5
6
7
= 12
Figure 7.32: Stability amplitude-frequency maps for sinusoidal actions. Eect of the slender-
ness ratio; = 0.925, p = 3.6, = /2.
Figure 7.32 shows maps corresponding to four values of the slenderness ratio . The graphics
were scaled so that the collapse boundary could be shown in the picture. As expected, an
increasing value of the slenderness ratio leads to a higher probability of collapse. However, it
7.6. Stability in the Amplitude-Frequency Range 133
can be realized that this growth is not linear, e.g. the dierence between the cases = 3 and
= 6 is clearly much larger than the cases = 9 and = 12.
Eect of the frequency parameter. The Scale eect
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
1
2
3
4
5
6
7
p = 1.0
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
1
2
3
4
5
6
7
p = 3.0
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
1
2
3
4
5
6
7
p = 5.0
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
1
2
3
4
5
6
7
p = 7.0
Figure 7.33: Stability amplitude-frequency maps for sinusoidal actions. Eect of the frequency
parameter; = 0.245, p = 3.6, = /2.
The next case examines the eect of the p parameter. Figures 7.33 address four contour maps
corresponding to increasing values of p. The results obtained from this plot are clear; the larger
the p values, the more unstable the RM dynamics gets. This fact conrms the so-called scale
7.6. Stability in the Amplitude-Frequency Range 134
eect pointed out by Housner [7]. He explained why the larger of two geometrically similar
blocks can survive forcing of a white-noise type, whereas the smaller block may topple. Since
for rectangular blocks, the expression for p depends solely on the size (R) of the block (see
Eqs. 2.5), large values of p correspond to smaller blocks. Therefore, from two blocks with the
same aspect ratio () and dierent size, the larger block is more stable than the smaller.
Eect of damping
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
1
2
3
4
5
6
= 0.80
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
1
2
3
4
5
6
= 0.90
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
1
2
3
4
5
6
= 0.95
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
1
2
3
4
5
6
= 0.99
Figure 7.34: Stability amplitude-frequency maps for sinusoidal actions. Eect of damping;
= 0.245, p = 3.6.
7.6. Stability in the Amplitude-Frequency Range 135
Finally, the eect of the damping parameter on the stability maps is computed through a
similar procedure. Figures 7.34 report contour stability maps for a varying value of the coe-
cient of restitution.
As in the preceding cases, there exist a diagonal boundary which divides the map into stable and
unstable regions; large values of damping correspond to more stable blocks. However, there is
an issue which must be highlighted for this case. As the value of increases (low damping),
the boundary between stable and unstable regions becomes wider. This allows that a number of
new orbit types (quasi-periodic, chaotic, etc) appear. As it will be examined in section 7.9 this
fact leads to a big wealth of dynamical behavior, including chaos.
Eect of the external phase
Finally, the eect of a change in the external phase is examined by considering the action as
a pure sine function ( = /2) or a pure cosine function ( = 0.0). Figure 7.35 report the
results of a variation in the external phase by an amount of /2 on the response of specimens 2,
3 and 4 with the experimental values of the parameters (Table E.1).
Roughly, a cosine-type function seems to provide a more stable behavior than a sine-type. This
is specially evident for specimen 4. The reason for this behaviour is that in the transient response
(which is the most important concerning stability), a sine-type function acts more as a pulse-
type than the cosine. However, further discussion on this subject will be provided in section 7.9
through a dierent perspective.
7.6. Stability in the Amplitude-Frequency Range 136
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
1 2 3 4 5 6
0.5
1
1.5
2
2.5
3
3.5
4
Specimen 2: = /2
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
1 2 3 4 5 6
0.5
1
1.5
2
2.5
3
3.5
4
Specimen 2: = 0.0
Collapse
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
1 2 3 4 5 6
0.5
1
1.5
2
2.5
3
3.5
4
Specimen 3: = /2
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
1 2 3 4 5 6
0.5
1
1.5
2
2.5
3
3.5
4
Specimen 3: = 0.0
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
0 5 10 15 20
2
3
4
5
6
7
8
9
10
Specimen 4: = /2
Collapse
No RM
m
a
x

|

|

(
d
e
g
)
PGA (m/s
2
)
f

(
H
z
)
5 10 15 20
2
3
4
5
6
7
8
9
10
Specimen 4: = 0.0
Figure 7.35: Stability amplitude-frequency maps for sinusoidal actions. Eect of the external
phase with the experimental values of the parameters (Table E.1)
7.7. Bifurcation Analysis 137
7.7 Bifurcation Analysis
A well known phenomenon of RM dynamics (see Hogan [46]) is the period cascade. A control
parameter is changed while the period of the orbit is monitored. Then, when certain values
of the control parameter are reached, the period "jumps" abruptly. This section deals with the
jumps of T/T
ext
for dierent control parameters, where "T" is the period of the orbit and "T
ext
"
is the external driving force period. A set of reference values for parameters and initial con-
ditions is chosen. Then, the non-dimensional system of Eqs. 5.50 is integrated for the varying
parameter while the rest of the parameters remain constant. The selected values for the analysis
are: = 0.245, = 0.925 and = 0.0 (cosine type), with initial conditions x
0
= 0.01, x
0
= 0.0
and
0
= 0.0.
It must be highlighted that, during the numerical integration, not all the orbits attained were
strictly periodic. As pointed out by many authors ([12],[14],[17],[39],[47]) dierent types of
orbits exist for RM dynamics in a given parameter range (section 7.9 deals with this matter).
The periods were obtained by choosing the mean Fourier frequency from the spectrum of the
response. Therefore, the periods for a non-periodic orbit (quasi-periodic or chaotic) represent
a rough approximation of the "periodic-equivalent" orbit. However, even within this approx-
imation, the results reported here have valuable interest to the overall understanding of RM
dynamics, as it will be shown.
Eect of amplitude
Figure 7.36 shows the values of the T/T
ext
period ratio for a varying external amplitude for
damped systems, and undamped motion. For damped motion, it can be seen from Fig. 7.36
a) that a well-dened T/T
ext
period ratio lies within the intervals: (2.7-3.3),(3.9-4.1),(4.9-
5.1),(6.5-6.7), etc. The central values of these intervals are: 3.0, 4.0, 5.0 and 6.5. As it will
be shown in section 7.9, if only bounded periodic motion was allowed, the T/T
ext
ratio would
follow an odd integer sequence; 1,3,5,7, etc. due to the presence of attractors (see section 7.9).
However, since for the computation of T/T
ext
, overturned orbits were also considered, the pe-
riod intervals do not exactly follow this odd integer sequence. On the other hand, it is evident
7.7. Bifurcation Analysis 138
1
2
3
4
5
6
7
0 2 4 6 8 10 12 14 16 18 20
T
/
T
e
x
t

=10
=12
=14
=16
=18
a) = 0.925.
1
2
3
4
5
6
7
8
9
10
0 2 4 6 8 10 12 14 16 18 20
T
/
T
e
x
t

=10
=12
=14
=16
=18
b) = 1.0.
Figure 7.36: Bifurcation. Amplitude scanning for dierent values and = 0.925, 1.0
that the "jumps" stand for all frequencies. However, not all these sub-harmonic gaps have the
same length (in terms of the control parameter ). For instance, the T/T
ext
= 4 only exists
between = 8 11, whereas T/T
ext=5
covers the range = 11 19.
When there is no damping (Fig 7.36 b)), although part of the shift structure still remains, the
jumps show a fuzzy random behaviour. This is due to the fact that, for = 1.0 (as it will be
analyzed in section 7.9), the predominant orbits are not periodic, but quasi-periodic and chaotic.
As a consequence, the period is not well dened resulting in this noisy appearance.
Eect of frequency
In Figure 7.37, the eect of the external frequency is monitored for various values and =
0.925, 1.0. From this picture, it can be seen that the predominant ratios: 3.0, 4.0, 5.0 and 6.5 are
still present. But the most important conclusion to be derived is that for certain values of the
external amplitude, orbits with a given T/T
ext
ratio only exist within a range of frequencies. This
result was reported already in [14]. There, sucient conditions for the existence of dierent
types of periodic orbits were derived. For other values of , these ranges, in general, change, as
shown in Fig. 7.37 b).
7.7. Bifurcation Analysis 139
2
4
6
8
10
12
4 6 8 10 12 14 16 18 20
T
/
T
e
x
t

=0.5
=1.0
=2.0
=3.0
=4.0
=5.0
=6.0
=7.0
=8.0
=9.0
=10.0
=11.0
=12.0
a) = 0.925
2
3
4
5
6
7
8
5 10 15 20 25 30 35
T
/
T
e
x
t

=5.0
=10.0
b) = 1.0
Figure 7.37: Bifurcation. scanning for dierent values
Eect of slenderness ratio
The eect of the slenderness ratio on the period shifts is monitored using the same procedure.
Figure 7.38 depicts the period ratio when the slenderness parameter varies. Three values of
were plotted for two xed values of = 4.0, 6.0 and = 10.0. For this case, the shift structure
1
2
3
4
5
6
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
T
/
T
e
x
t
(rad)
=0.80
=0.925
=1.0
= 4.0
1
2
3
4
5
6
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
T
/
T
e
x
t
(rad)
=0.80
=0.925
=1.0
= 6.0
Figure 7.38: Bifurcation plot for scanning for two values of amplitude, = 10.0 and =
0.8, 0.925, 1.0
7.7. Bifurcation Analysis 140
is still present. When damping is allowed, the T/T
ext
period ratio lies close to 3.0. On the other
hand, undamped motion oscillates between the modes T/T
ext
= 3.0 and T/T
ext
= 5.0. However,
as it can be observed from Fig.7.38 this periodicity changes when the amplitude varies.
Eect of damping
Finally, Figs. 7.39 a) and 7.39 b) depict the bifurcation graph of the period ratio when the
parameter varies. In this case, the external frequency was xed to = 10.0 and six dierent
levels of external amplitude were examined. Figure 7.39 a) shows the graphs for low values of
whereas Fig. 7.39 b) plots T/T
ext
for large external amplitudes. First, it is noticed that the
transitions (jumps) for large values are larger than those corresponding to a lower external
amplitude action. On the other hand, the ratios do not follow the integer scheme, which is
particularly evident for low amplitudes. The author is unable to provide an explanation of why
this happens.
1.5
2
2.5
3
3.5
0.8 0.82 0.84 0.86 0.88 0.9 0.92 0.94 0.96 0.98
T
/
T
e
x
t

=3.0
=2.0
=1.0
=3.0
=2.0
=1.0
a) = 1 3
2.5
3
3.5
4
4.5
0.8 0.82 0.84 0.86 0.88 0.9 0.92 0.94 0.96 0.98
T
/
T
e
x
t

=5.0
=7.0 =9.0
b) = 3 7
Figure 7.39: Bifurcation plot for scanning. = 10.0
There are two main limitations in the study presented in this section. First, as stressed above,
not all the orbits were periodic and, hence, the computed periods by means of the response
spectrum give only an approximation to the real period. And, second, the characteristics of the
gures shown could change when other set of initial conditions is selected. However, the main
7.8. Delay-coordinates Analysis 141
eort of this analysis was not quantitative but qualitative. It consisted on showing the period
jumps as an inherent feature of RM dynamics. In order to study the behavior considering the
whole range of initial conditions a new technique is implemented in the next section.
7.8 Delay-coordinates Analysis
This section addresses the so called delay-coordinates analysis. The method of delays is a well-
known phase-space reconstruction technique [65],[66]. Vectors in the embedding space (see
[66]) are formed from time delayed values of the scalar measurements:
s
n
= (s
n(m1)d
, s
n(m2)d
, ..., s
n
) (7.16)
where m is the so-called embedding dimension and d is the delay time. For this study, an
embedding dimension of two is used. From the values of the time-series, a two-column le is
generated with value ((), ( d)).
The delay-coordinate method is used as the rst approach to detect modes and correlations at
dierent time scales. The main orbital types described in section 6.4 are analyzed here through
the delay method. In section 7.9 the phase space is fully investigated through other tools.
Reconstruction of modes for harmonic forcing regime
Figure 7.40 shows the delay plots for periodic (1,1-7) Modes under harmonic forcing for various
delay lags. It is noticed that as the delay is shifted, the modes rotate around the origin (0.0).
On the other hand Fig. 7.41 shows a sequence of the delay plots of the periodic (1,3) Mode for
increasing steps of delay. The basic (1,3) Mode structure, already found in Fig.6.28 appears for
d = 0.1s. As the delay lag increases, the shape is deformed and rotated until d = 0.8s. At this
time, the mode appears again but rotated with an angle of . The other modes follow a similar
trend.
Figure 7.42 shows the delay plots for various lags corresponding to two quasi-periodic responses
obtained for harmonic motion with constant amplitude. It is noticed that the regularity present
7.8. Delay-coordinates Analysis 142
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
X
1
(
t
-
d
)
X1(t)
0.1s
0.5s
1s
2s
(1,1) Mode
-0.2
-0.1
0
0.1
0.2
-0.2 -0.1 0 0.1 0.2
X
1
(
t
-
d
)
X1(t)
0.1s
0.5s
1.0s
(1,3) Mode
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
-0.3 -0.2 -0.1 0 0.1 0.2 0.3
X
1
(
t
-
d
)
X1(t)
0.1s
0.5s
1.0s
2.0s
(1,5) Mode
-0.4
-0.2
0
0.2
0.4
-0.4 -0.2 0 0.2 0.4
X
1
(
t
-
d
)
X1(t)
0.1s
0.5s
0.8s
1.2s
2.5s
(1,7) Mode
Figure 7.40: Delay plots for periodic modes under harmonic forcing. Numerical simulation
for periodic modes no longer holds. These structures appear as containing all the periodic
modes in competition.
7.8. Delay-coordinates Analysis 143
Figure 7.41: Delay plots of periodic (1,3) Mode for dierent time lags
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
X
1
(
t
-
d
)
X1(t)
0.1s
0.5s
1.0s
2.0s
Quasi-periodic mode I
-0.4
-0.2
0
0.2
0.4
-0.4 -0.2 0 0.2 0.4
X
1
(
t
-
d
)
X1(t)
0.1s
0.5s
1.0s
2.0s
Quasi-periodic mode II
Figure 7.42: Delay plots for quasi-periodic modes under harmonics forcing with constant am-
plitude. Numerical simulation
Figure 7.43 plots the delay plot corresponding to a chaotic orbit generated with a constant-
amplitude harmonic load. For this case, no pattern is recognized. The orbits are fully chaotic
and the only symmetry which is maintained is the reections about the principal diagonal.
7.9. Poincar Surface of Section Analysis 144
-0.4
-0.2
0
0.2
0.4
-0.4 -0.2 0 0.2 0.4
X
1
(
t
-
d
)
X1(t)
0.1s
0.5s
1.0s
2.0s
Figure 7.43: Delay plot for chaotic orbit under harmonic forcing with constant amplitude. Nu-
merical simulation
7.9 Poincar Surface of Section Analysis
The analysis of phase space trajectories is a basic concept of nonlinear analysis. Even when no
analytical results are possible, the visual inspection of the phase plane trajectories gives hints
about the dynamics, i.e. the attractor of the system. Since the dimension of the phase space
is very large, Poincar introduced a dimension reduction technique, namely, the Poincar Sur-
face of Section (PSS), which allows both the visualization and quantication of the dynamics.
This section presents the results of an extensive numerical study over the parameter range by
systematic use of the PSS.
PSS and stability
Many authors have used the PSS method to nd existence and stability boundaries in the
amplitude-frequency space ([9],[12],[17],[34],[47]). However, the majority of them focused
their studies on the stability of orbits integrated with quiescent initial conditions.
On the other hand, Hogan [12] has pointed out the necessity of determining stability boundaries
not only from initial conditions but also for a range of them. Recently, Rega & Lenci [67] has
used the PSS procedure for a cluster of initial conditions.
7.9. Poincar Surface of Section Analysis 145
In this section, a numerical study over the initial conditions and parameter values is accom-
plished through intensive use of the Poincare Surface of Section (PSS) method. First, the phase
space structure is investigated under the assumption of an harmonic load by means of an strobo-
scopic PSS. Then, the eect of white noise on such structure is investigated for dierent values
of noise strength. Finally, synthetic earthquakes are used as input signals.
Critical points and phase space structure
From the system of Eqs. 5.50 it is found:
= v
v = 1 + G(t, , ) + a
g,
(7.17)
where the function G(t, , ) =

[

+ ln()/ ] has been introduced and a


g,
= a
g
cos()
represents the coupling between the external action a
g
(t) and the phase . If G is set to zero, no
barrier (and, hence no oscillatory motion) and damping eects are possible. From Eqs. 7.17 it
can be found that for undamped-unforced systems there are two xed points: (0, 0) and (1, 0).
Since = ||/, in the plane (Re(), Re(

)) the point (1, 0) unfolds into the points (1, 0) and


(1, 0). These unstable equilibrium positions corresponds to the block standing on one edge.
The origin (0, 0) is an stable singular point while the points (1, 0) and (1, 0) are unstable
hyperbolic saddle points. Therefore, the classication of critical points for the unperturbed
phase space is as follows.
1. (0,0) stable point,
2. (1,0) unstable saddle point,
3. (-1,0) unstable saddle point.
This conguration divides the phase space into stable (I) and unstable (II) regions (see Fig. 7.44).
Trajectories near the stable singular point remain in its neighborhood, while trajectories near the
hyperbolic singular points will diverge from them. The locus is given through the equations:
v = 1
v = 1 +
(7.18)
7.9. Poincar Surface of Section Analysis 146
This closed curve which connects the two unstable hyperbolic points forms the so-called hete-
roclinc connection. Figure 7.44 depicts some orbits in the phase plane (Re(

/), Re(/)) for


the unperturbed phase space. The separatrix given by Eq. 7.18 is plotted in thick black line.
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
R
e
(

p
)
Re(/)
(I)
(II)
W
a
s
W
a
s
W
a
u
W
a
u
W
b
u
W
b
u
W
b
s
W
b
s
Figure 7.44: Singular points in the phase plane for undamped, unforced systems.
Additionally, the ow directions are indicated through arrows. Stable W
s
and unstable W
u
man-
ifolds [68] are also indicated in order to highlight the hyperbolic nature of the points (1,0) and
(-1,0).
It is a well known fact from chaos theory (see [68] or [69]) that when the system is forced, the
original structure is perturbed and, hence, heteroclinic intersections between the stable and un-
stable manifolds may occur. Conditions for this intersections can be found analytically through
the Melnikov method ([13],[15],[67]). However, this study is focused on a full numerical explo-
ration over the phase-space, where the main eort is to nd a possible quantication of chaos
that can be useful to earthquake engineering.
7.10. PSS for Harmonic Load 147
7.10 PSS for Harmonic Load
Denition of the Poincar surface of section (PSS)
By recovering the system of Eqs. 5.50 the following transformation is performed on the state
variables (
1
,
2
,
3
,
4
):
X1(
1
,
2
,
3
,
4
) =
1
cos(
3
)
X2(
1
,
2
,
3
,
4
) =
2
cos(
3
)
X3(
1
,
2
,
3
,
4
) =
4
modulo 2
(7.19)
The reason for the denition of
4
is that this variable is unbounded since it is proportional to
t. However,
4
only occurs as the argument of a cosine, and hence it can be regarded as an
angle (see Fig. 7.45). Thus, the phase space coordinates can be taken as X1, X2, X3, where
Figure 7.45: Denition of Stroboscopic Poincar Surface of Section.
X3
4
modulo 2. The new variables are, in the RM limit:
X1 = Re(/)
X2 = Re(

/)
X3 = t modulo T
ext
(7.20)
As it is usual for forced systems under harmonic pulses, the PSS is dened by the following
Stroboscopic map on the plane (X1, X2):
X3 = 0 t = T
ext
(7.21)
7.10. PSS for Harmonic Load 148
The PSS reduces the original three-dimensional problem in the state space (X1, X2, X3) to a
two-dimensional problem in the plane (X1, X2). The system of Eqs. 5.50 is integrated over a
grid of initial conditions generated randomly. When the PSS condition of Eq. 7.21 is satised,
a Poincar point is plotted in the reduced phase plane (X1, X2).
Main types of PSS and their associated orbits
For each of the main types of orbits of the state space (periodic, quasi-periodic, chaotic, chaotic
and overturning) there is a corresponding PSS. Periodic orbits correspond to a nite number of
PSS points. Figure 7.46 depicts the main periodic modes found during the numerical investiga-
tion. As highlighted in [46], for = 1.0 periodic subharmonic orbits are neutrally stable and,
hence, they are unlikely to be seen in numerical computations. This fact has been conrmed in
the present study. However, some undamped subharmonic orbits have been found by straight-
forward integration. Figure 7.47 shows some of those unstable subharmonic modes. although,
as stressed above, they are unstable and they vanish after integration over a short number of
periods (typically, 10 or 20). As Hogan shown [12], no even symmetric modes are possible.
Those modes were not found after an extensive numerical investigation.
Quasi-periodicity
The next type of bounded rocking response was rst reported by Yim & Lin [47]. It is character-
istic of systems with two or more incommensurate competing frequencies. The two candidate
frequencies are the external driven force frequency and the internal frequency. However, only
for = 1.0, the internal frequency is a well-dened quantity during. In this case, the frequency
is given by the Housner frequency (Eq. 6.4). Therefore quasi-periodicity only occurs for the
case = 1.0.
The PSS of quasi-periodic orbits have the appearance of islands. After one time around the
large circumference of the torus (Fig. 7.45), the Poincar point spins around the small torus
circle resulting in small circles in the PSS. Typical PSS of quasi-periodic structures are shown
in Fig. 7.48.
7.10. PSS for Harmonic Load 149
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
-0.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25
X
2
X1
a) Symmetric (1,1) Mode
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
-0.15 -0.1 -0.05 0 0.05 0.1 0.15
X
2
X1
b) Symmetric (1,3) Mode
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2
X
2
X1
c) Symmetric (1,5) Mode
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
-0.3 -0.2 -0.1 0 0.1 0.2 0.3
X
2
X1
d) Symmetric (1,7) Mode
Figure 7.46: Periodic modes in the reduced (X1, X2) state space
As it was highlighted by Hogan [12], for a given set of parameters, many dierent modes coex-
ist if the initial conditions are free to change. These coexistence of modes is not observable if
only quiescent initial conditions are chosen. In Fig. 7.48 three types of stability island structures
(3, 5 and 7) are depicted. The computation was performed by increasing forward the value of
X2 while X1 remained xed for the same set of parameters (, , , ).
Along the numerical computation it has been observed that this island structures appear in odd
integer sequence: 1, 3, 5, 7, etc. This sequence exhibits periodicity; the island structures appear
and vanish as long as a control parameter is changed.
7.10. PSS for Harmonic Load 150
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2
X
2
X1
a) Unstable symmetric (1,5) Mode
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
X
2
X1
b) Unstable symmetric (1,7) Mode
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
X
2
X1
a) Unstable asymmetric (1,8) Mode
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
-0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2
X
2
X1
b) Unstable asymmetric (3,9) Mode
Figure 7.47: Unstable periodic modes in the reduced (X1, X2) state space
On the other hand, it has been found that no stable periodic orbits for undamped systems ex-
ist; only quasi-periodic and chaotic behavior is found. Although periodic modes, for = 1.0,
were observed during the numerical exploration, they were unstable and they vanished after
some integration steps. On the other hand, damped systems lack of quasi-periodic behavior and
only periodic or overturning response exist. These results are in agreement with previous works
[12],[15],[46].
7.10. PSS for Harmonic Load 151
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
X
2
X1
X2=0.0405
X2=0.1457
X2=0.4457
Figure 7.48: Poincar section for quasi-periodic modes
Spanos & Koh [14] studied the stability of periodic modes. As highlighted in [46] the periodic
modes can be unstable depending upon which parameter values are chosen. In the following,
the eect of parameter changes on the phase-space structure is addressed. The reference values
assumed for the parameters are: = 0.245, = 1.0, = 8.0, = 10.0.
Eects of damping
Hogan [12] has pointed out that the phase-space structure of undamped systems remains when
damping is present. On the other hand, due to damping, the phase space shrinks and volumes
are not preserved under the ux of the equations of motion. It is therefore relevant to analyze
how the damping aects the phase space.
The system of Eqs. 5.50 can be recast into vector form as:

=

F() (7.22)
If V denotes the phase space volume, by the divergence theorem it is found that:
dV
dt
=
_
V
d
4
(



F) (7.23)
where
_
V
signies the integral over the phase-space volume V. From Eqs. 5.50, it is found that:



F =
ln(
2
)l

2
1
(7.24)
7.10. PSS for Harmonic Load 152
and, hence, Eq. 7.23 reads:
dV
dt
=
ln(
2
)

_
v
d
4
l

2
1
(7.25)
Now, by introducing the expression of

= l/
2
1
in the RM limit (see appendix C), the following
equation holds:
dV
dt
= ln(
2
)

j
(t t
j
)V (7.26)
where, as in the appendix Cthe sumof deltas is extended over the impact times t
j
. By integrating
through an impact time it is found:
V = V
0

2
(7.27)
where V
0
represents the initial phase-space volume. Therefore, at every impact there is a reduc-
tion of the phase space volume by an amount of
2
, equal to the energy reduction (Eq. 2.16).
This eect can be veried in the PSS shown in Fig. 7.49.
On the other hand, when damping is present, the quasi-periodic structure is destroyed and trans-
formed into periodic. For instance, it has been noticed that moderate values of damping destroy
the island structure. As the damping decreases, a 5-island structure is observed (Fig. 7.49 c) and
d)). From Fig. 7.49 it is observed that as long as damping values increase, the island structures
a) = 0.70 b) = 0.85
c) = 0.98 d) = 0.99
Figure 7.49: PSS. Eect of damping
are destroyed. Those modes collapse into a single big attractor at the origin.
7.10. PSS for Harmonic Load 153
On the other hand, it is observed from Fig. 7.49 that there are accumulation of points following
a wave pattern in the rst and third quadrant of the gure. This phenomenon is particularly
evident for large values.
The eect is well known in chaos theory and it is due to the so called heteroclinic intersections
between the stable and unstable manifolds [68]. In the context of RM dynamics it has been only
recently investigated [67]. As stated above, these intersections only occur in the rst and third
quadrants, which clearly evidences a lack of symmetry in the dynamics. This lack of symmetry
was fully explained by Hogan [13] through analytically extending the stable and instable man-
ifolds and then by considering the relevant intersections. However, the theoretical justication
of this fact lies out of the scope of the present work.
Eect of amplitude
In an intensive numerical study, a set of 100 PSS was generated ranging from = 0.1 to
= 10.0 with an step of "0.1". Figures 7.50 and 7.51 show the results for the undamped regime
for dierent values. With no damping at hand, as expected, no periodic structure is observed;
a) = 0.1 b) = 0.5
b) = 1.0 b) = 2.0
Figure 7.50: PSS. Eect of external amplitude for undamped motion I. Parameters: = 0.245,
= 10.0, = 0.0
only chaotic and quasiperiodic patterns are found.
7.10. PSS for Harmonic Load 154
b) = 3.8 b) = 6.0
c) = 8.0 d) = 10.0
Figure 7.51: PSS. Eect of external amplitude for undamped motion II. Parameters: = 0.245,
= 10.0, = 0.0
It is also observed that as the amplitude increases, the manifold intersection amplies. Starting
with all the points within the regular stable region, the points start to accumulate in wave patterns
at quadrants rst and third as the amplitude increases.
As stated above, the attractors show periodicity. For instance, the 5-island structure appears
once a value of = 3.8 is reached. It then persists until the value = 6.0 at which the structure
vanishes. For damped systems, odd n values of subharmonic (1, n) modes coexist at dierent
external amplitude levels. Figures 7.52 and 7.53 show how the (1,3), (1,5) and (1,7) periodic
modes emerge and vanish when is ranged from 1.0 to 8.0.
Eect of frequency
The eect of the external frequency on the PSS for undamped systems is shown in Figs. 7.54
and 7.55. When no damping is present, a single big attractor at the origin with a complex
structure is observed (Fig. 7.54). The number of PSS points is reduced for low values since
collapse is likely to occur. As increases, the 5-island quasi-periodic structure is found if
> 10 (Fig. 7.55).
7.10. PSS for Harmonic Load 155
a) = 1.0 b) = 2.0
c) = 3.0 d) = 4.0
Figure 7.52: PSS. Eect of external amplitude for damped motion I. Parameters: = 0.925
= 0.245, = 10.0, = 0.0
e) = 5.0 f) = 6.0
g) = 7.0 h) = 8.0
Figure 7.53: PSS. Eect of external amplitude for damped motion II. Parameters: = 0.925
= 0.245, = 10.0, = 0.0
On the other hand, the unstable manifold oscillations around the stable manifold become more
tight as increases. When damping is allowed, the solutions of the systems cannot ll a volume
of phase space, since dissipation is synonymous with a contraction of volume elements under
7.10. PSS for Harmonic Load 156
a) = 4.0 b) = 5.0
c) = 6.0 d) = 7.0
Figure 7.54: PSS. Eect of external frequency for undamped motion I
the action of the equations of motion. As stressed above for the case of amplitude scanning,
e) = 8.0 f) = 9.0
g) = 10.0 h) = 12.0
Figure 7.55: Pss. Eect of external frequency for undamped motion II
the "island" structure vanishes and only periodic and chaotic behavior exists. Figures 7.56 and
7.57 show the Poincar for dierent values of and a damping of = 0.925. From comparison
between Figs. 7.54, 7.55 and 7.56, 7.57 it can be noticed that much of the structure for undamped
motion persists when damped is allowed. For small values of the external frequency (typically
7.10. PSS for Harmonic Load 157
a) = 4.0 b) = 5.0
c) = 6.0 d) = 7.0
Figure 7.56: PSS. Eect of external frequency for damped systems I
e) = 8.0 f) = 9.0
g) = 10.0 h) = 12.0
Figure 7.57: Pss. Eect of external frequency for damped systems II
< 6) the single attractor observed for undamped motion is still found. For higher values of
, the 3 and 5 structures are still present. However, in this case, those structures are not islands
(as for undamped motion) but attractors.
The fractal structure of the attractor at the origin must be highlighted. The fractal nature of this
attractor was observed by Hogan [12] and Sinopoli [34].
7.10. PSS for Harmonic Load 158
Eect of slenderness ratio
For undamped motion, it is noticed that quasiperiodicty only appears for certain values of .
Figure 7.58 shows the Poincar sections for six dierent values of the slenderness ratio. It is
a) = 12 b) = 9
c) = 6 d) = 4
e) = 3 f) = 2
Figure 7.58: PSS. Eect of slenderness ratio for undamped systems
noticed that as decreases (the block becomes less slender), more quasi-periodic modes coexist.
For instance, the complete 1, 3, 5, 7 island sequence is achieved only for = 2.
On the other hand, the stable-unstable manifold intersections do not change substantially when
is varied. If damping is present (Fig. 7.59) it is found that the 5-island structure (now attractors)
persists for a wide range of slenderness ratio values. Heteroclinic intersections are, as in the
undamped case, apparently independent of .
7.10. PSS for Harmonic Load 159
a) = 12 b) = 9
c) = 6 d) = 4
e) = 3 f) = 2
Figure 7.59: PSS. Eect of slenderness ratio for damped systems
Eect of external phase
The external phase is changed by an amount of /2 as in the stability maps in order to
analyze the eect of a sinusoidal action on the PSS. Figure 7.60 shows the PSS for a sinusoidal
Figure 7.60: PSS. Eect of the external phase for undamped systems
action for = 0.245, = 1.0, = 8.0, = 10.0 (compare with Fig. 7.51 c)). It is observed that
7.11. Eect of White Noise on the PSS 160
there is a shift in the velocity axis (X2) which displaces the PSS in the vertical direction.
On the other hand, from Figs. 7.61 a) and b) it is shown that the odd integer 1, 3, 5,.. subhar-
monic structure is still present for sinusoidal actions.
a) = 4 b) = 8
Figure 7.61: PSS. Eect of the external phase for damped systems
7.11 Eect of White Noise on the PSS
As a further step towards the understanding of stability of RM under random loading, the eect
of a sinusoidal action perturbed with white noise is examined next. As it will be shown, the
noise promotes transitions between the attractors. The aim of this section is to test how the
phase space structure is aected by an external noise. Figure 7.62 shows a typical input load
consisting on the superposition of a sinusoidal function and a white noise signal of a given noise
strength.
Eect of white noise on orbits
Along the numerical computation it has been observed that the noise does not aect in the same
proportion all the orbits. Since the PSS is constructed as stroboscopic shots of trajectories in
the extended phase-space (see Fig. 7.45), it is relevant to investigate how these trajectories are
modied by dierent levels of noise. Figure 7.63 presents the results of the computation for the
subharmonic (1,3) Mode with noise levels ranging from 10% to 100% of the external amplitude.
It can be noticed that the form persists even for high values of noise. The results for the mode
(1,1) are analogous.
7.11. Eect of White Noise on the PSS 161
Figure 7.62: Typical input load time series for a white noise test
-15
-10
-5
0
5
10
15
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
X
2
X1
a) 10% noise
-15
-10
-5
0
5
10
15
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
X
2
X1
b) 30% noise
-15
-10
-5
0
5
10
15
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
X
2
X1
c) 50% noise
-15
-10
-5
0
5
10
15
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3
X
2
X1
d) 100% noise
Figure 7.63: Eect of white noise on subharmonic (1,3) Mode
On the other hand, quasi-periodic and chaotic orbits are much more sensitive to noise. This
suggests that periodic modes (without considering undamped-unstable modes) are more robust
to noise than quasi-periodic and chaotic orbits. Figure 7.64 presents the phase space trajectories
for the chaotic orbit of Fig.6.31. It can be noticed that at maximum level of noise (100%),
7.11. Eect of White Noise on the PSS 162
-20
-15
-10
-5
0
5
10
15
20
-10 -8 -6 -4 -2 0 2 4 6 8 10
X
2
X1
a) 10% noise
-25
-20
-15
-10
-5
0
5
10
15
20
-10 -5 0 5 10 15
X
2
X1
b) 30% noise
-25
-20
-15
-10
-5
0
5
10
15
20
-8 -6 -4 -2 0 2 4 6 8 10
X
2
X1
c) 50% noise
-20
-15
-10
-5
0
5
10
15
20
-6 -4 -2 0 2 4 6
X
2
X1
d) 100% noise
Figure 7.64: Eect of white noise on sinusoidal input load for a chaotic orbit
the orbit contained the (1,1) and (1,3) modes. In this sense, it is observed that noise is able to
promote transitions between modes.
Eect of white noise on the PSS
Now, the eect of white noise on the PSS is examined. For this computation, the parameters
( = 0.245, = 0.925, = 10.0, = 12.0) where chosen, Then a cosine-type signal was
used to obtain the orbits by forward integrating from a cluster of initial conditions. Finally, the
stroboscopic PSS is found. Figure 7.65 shows the PSS for moderate (10%) and high (100%)
noise levels. It can be noticed that when moderate levels of noise are added to the harmonic input
signal, the structures in the PSS are partially destroyed. In Fig. 7.65 a) no attractor sequences
are shown besides the single attractor at (0.0). On the other hand if high levels of noise are
allowed, the phase space patterns are fully destroyed, as it is shown in Fig. 7.65 b). However, it
can be concluded that the single attractor at the origin persists even for high values of noise.
7.12. PSS for Random Loading 163
a) 10% noise b) 100% noise
Figure 7.65: Stroboscopic PSS for sinusoidal action with = 10.0 and = 12.0 on a block
with = 4, p = 3.6s
1
and = 0.925
7.12 PSS for Random Loading
PSS for random loading. Stroboscopic PSS
The structure of the phase space is now analyzed for earthquake actions. For this case, the
output variables are (Re(), Re(

)) = (X1, pX2). This map is the dimensional homologous


to the non-dimensional (X1, X2), introduced for harmonic motion. It must be stated, that for
the dimensional variables, the separatrix given in Eq. 7.18 is now rescaled in units along the
horizontal direction and in p units along the vertical direction. Figure 7.66 shows the separa-
trix (thick red line) of the stroboscopic PSS along with the Poincar points for a typical random
loading computation. The vertices of the rhombus are now located at and p.
Figure 7.66: Denition of separatrix for PSS under random loading
7.12. PSS for Random Loading 164
On the other hand, when parameters are varied, there are changes in the dispersion of the
Poincar points from the stable domain outward from the separatrix. As it will be discussed
in chapter 8, the collapse phenomenon can be understood as the probability of crossing this
separatrix. Therefore, in the following, this eect is studied through numerical computation for
several ranges of the system parameters and the external PGA.
Eect of damping
As shown above in this section, damping shrinks the phase space. This eect is shown again in
Fig. 7.67 for random loading. When there is no damping (Fig.7.67 a)), the whole stable region
a) = 1.0 b) = 0.9
a) = 0.7 b) = 0.5
Figure 7.67: PSS for random loading. Eect of damping, = 0.245, p=3.6
within the separatrix is dense of Poincar points. As damping increases, the points collapse to
the attractor at (0,0).
Eect of PGA
Figure 7.68 shows the PSS of earthquake 20 for dierent amplication factors. As the external
load increases, the Poincar points scatter over the (X1, X2) plane. This eect can be thought
as the opposite to the damping shrinkage.
7.12. PSS for Random Loading 165
a) PGA=2.7m/s
2
b) PGA=5.4m/s
2
c) PGA=8.1m/s
2
d) PGA=10.8m/s
2
e) PGA=16.2m/s
2
f) PGA=27.0m/s
2
Figure 7.68: PSS for random loading. Eect of PGA, = 0.245, p = 3.6, = 0.925
Eect of slenderness ratio
Figure 7.69 depicts four PSS corresponding to blocks with dierent aspect ratio. It is noticed
that slender blocks have more dense attractors. This fact is due to the rescaling of the phase
plane when the nondimensional variables (X1, X2) are translated to the dimensional variables
(Re(), Re(

)). There is an overall scaling factor of = arctan(1/). For small angles tan
and, hence 1/. Therefore, if increases, there is a global resizing in the (Re(), Re(

))
plane by an amount of 1/. Furthermore, the separatrix is rescaled in the same proportion. This
fact has a dramatical eect on the stability since, when a cluster of equally distributed initial
conditions is given, the probability of falling within the stable domain decreases rapidly with .
7.12. PSS for Random Loading 166
a) = 3 b) = 6
c) = 9 d) = 12
Figure 7.69: PSS for random loading. Eect of slenderness, p = 3.6, = 0.925.
Eect of the frequency parameter
The frequency parameter p has a notable eect on the shape of the attractor and, hence, the
stability of the system. Figure 7.70 shows the PSS obtained with synthetic earthquake 20 with
PGA=2.7m/s
2
and = 0.245, = 0.925 for the p values of 1, 2, 4 and 6. Since the dimensional
velocities Re(

) are equal to pX2, there is a shrinkage along the vertical direction in the
(Re(), Re(

)) plane as p varies. As a consequence, the attractor becomes more narrowand large


in the vertical direction as p increases. Additionally, the attractor rotates counter-clockwise as
p increases.
Threshold crossing distribution
As stated above, as issue which is directly related with collapse mechanism is the rate of
Poincar points which cross the separatrix. Figure 7.71 depicts the results of this computa-
tion when the value of the external PGA (normalized by the West acceleration) is varied. The
number of crossing points is normalized to the total amount of initial points. It can be noticed
that the so-called diusion mechanism increases smoothly until PGA a
w
. Once the West
acceleration is reached, the crossing rate of the separatrix increase signicantly.
7.12. PSS for Random Loading 167
a) p = 1 b) p = 2
c) p = 4 d) p = 6
Figure 7.70: Two types of stroboscopic PSS for random loading
1.45
1.5
1.55
1.6
1.65
1.7
1.75
1.8
1.85
1.9
1.95
0 0.5 1 1.5 2 2.5
N
u
m
.
p
o
i
n
t
s
/
N
PGA/a
w
Figure 7.71: Number of Poincar points crossing the separatrix as function of the external PGA
normalized with the West acceleration
Attractor structure
In the computations given above, it has been observed that the attractor structure consists on
three main parts (Fig. 7.72):
1. Central dense region.
2. Surrounding scattered region.
3. Two diagonal wings (or arms).
7.12. PSS for Random Loading 168
As it has been noticed, the transitions from the dense region to the scattered domain can be
more smooth or abrupt depending on the system parameter values. On the other hand, the
length and orientation of the wings depends strongly on the frequency parameter p. Figure. 7.72
depicts these three main parts of a typical attractor under random loading. As a rst approach
Figure 7.72: Analysis of attractor structure. Main regions
to understand the distribution of points in the attractor, a simple Gaussian probability density
function is proposed. The starting point is joint probability distribution (, ) in the non-
dimensional variables (, ) = (r/, r/). This is assumed to be a normal distribution with
zero mean and standard variation sigma

. Since in the new phase space, the variables are:


X1 = , X2 = p , when the new probability distribution function (X1, X2) is derived, two
eects must be considered:
1. Dilatation in both axes by an amount of .
2. Dilatation in the vertical axis by an amount of p.
The resulting distribution function is given by:
(X1, X2) =
1

2
exp (
1
2
2
(X1
2
+ X2
2
/p
2
)) (7.28)
where the standard deviation has to be tted according to the system parameters (, p,) and
the external PGA.
7.13. Recurrence Plots and Dynamics 169
Figure 7.73 a) plots the probability distribution of Eq. 7.28 along with various contour lines
in the (X1, X2) plane. On the other hand, Fig. 7.73 b) compares the probability distribution of
-10
-5
0
5
10
-10
-5
0
5
10
p(X1,X2)
0.4
0.3
0.2
0.1
X1
X2
a)
0
0.05
0.1
0.15
0.2
-20 -15 -10 -5 0 5 10 15 20
F
r
e
q
.

0
ARM
FIT: Gaussian
b)
Figure 7.73: Probability distribution function for random loading in the phase space: a) Surface
plot of (X1, X2) from Eq. 7.28, b) comparison between Eq. 7.28 and the distribution of X1 in
the attractor
Eq.7.28 and the histogram for the distribution of X1 in the attractor for a xed value of X2.
As it is indicated in the gure (Fig. 7.73 b)), the simple density function proposed is not able
to account for the eect of the wings. However, it may be still used as a rst estimate of the
distribution of Poincar points in the attractor.
7.13 Recurrence Plots and Dynamics
Recurrence plots (RP) are a graphical tool for the study of nonlinear dynamical systems in a
manner which is apparent to the eye. RPs can highlight periodicity, unstable periodic orbits,
and other characteristics of the underlying dynamics. RP were rst introduced by Eckmann and
Ruelle [70] in 1987. Since then they have been used recently in many dierent areas in order to
nd out the dynamical structure of the system under study [65].
The idea underlying the RP is extremely simple. From the starting time series, t
i
, x
i
the distance
of every two points (x
i
, x
j
) is scanned. Then, the pixel lying at (i,j) is black if the distance falls
within a specied threshold corridor and white otherwise.
An attractor is the open set of densely embedded unstable periodic orbits (UPO) of all periods.
7.13. Recurrence Plots and Dynamics 170
Therefore, the determination of UPOS as the basic pieces of an attractor provide useful infor-
mation about the dynamics of the structure.
General qualitative interpretation can be derived from a given RP [65]. The main characteristics
are:
1. White zones in the plot are due to abrupt changes in the system dynamics.
2. A diagonal line occurs when the trajectory visits the same region of the embedded space
several times. Processes with stochastic behavior cause none or very short diagonal lines,
whereas deterministic processes cause longer diagonals and less single, isolated recur-
rence points.
3. Periodic patterns indicate cyclicities in the process; the time distance between periodic
patterns corresponds to the period.
4. Trends and drifts of the system appear in the RP as fading to the upper left and lower right
corners
5. Diagonal lines parallel to the main diagonal suggest determinism since the evolution of
states is similar at dierent times. If these lines are periodic, the unstable periodic orbits
may be present.
6. Vertical and horizontal lines reect the fact that some states do not change or change
slowly for some time. The so called laminar states are present.
Several parameters can be used in order to estimate the RP. The principal ones are:
The embedding dimension m.
The threshold corridor .
The delay time d.
One chooses the threshold corridor value in order to make the structure of the RP apparent. The
lower the value, the more accurate are the results obtained. However, a very low value of
will shade the RM and the structure will be lost. Other relevant parameter is the topological
norm. In this work, the simplest Euclidean norm was used since it provided results adequate to
the problem.
7.13. Recurrence Plots and Dynamics 171
On the other hand, several embedding dimensions and delay times needed to be tested for
various RP. Figure 7.74 shows the RP for embedding dimensions of 20, 10, 5 and 1 of ()
under random loading. It is clear from Fig. 7.74 that the structure is better highlighted in the
m=20 m=10
m=5 m=1
Figure 7.74: Recurrence plot of () under random loading for four dierent embedding dimen-
sions
RP of m=1, i.e. no embedding.
On the other hand, the delay is monitored for dierent values of m > 1. Figure 7.75 shows RP
of delay times of 10 and 1 for m = 10 and = 0.1. The value of d = 1 was found as the most
d=10 d=1
Figure 7.75: Recurrence plot of () under random loading for two dierent delay times
suitable choice in order to highlight the structure of the dynamics.
7.13. Recurrence Plots and Dynamics 172
Harmonic forcing
The recurrence plots corresponding to the main bounded orbital types under harmonic forced
regime, namely, periodic, quasi-periodic and chaotic are computed and discussed. In every
graph, the corresponding time series has also been included to better illustrate the patterns
present in the RP.
Periodic modes
The RP corresponding to the most common periodic mode, the (1,1) Mode, is depicted in
Fig. 7.76. The RP of this mode consists on a chessboard type pattern whose basic structure
-0.2
-0.1
0
0.1
0.2
0 2 4 6 8 10
R
e
(

)
Time (s)
Figure 7.76: Recurrence plot of stable periodic (1,1) Mode. Embedding dimension=1, delay=1,
threshold corridors of 0.1
is an oval. The complete periodicity results in a homogeneous graph. The distance between the
ovals is the period of the time series.
Subharmonic modes generate RPs where the basic oval structure of the (1,1) Mode is deformed
due to the presence of subharmonic frequencies. Figure 7.77 plots the RP graph corresponding
to the subharmonic (1,3) Mode The distance between white ovals (see Fig. 7.78) corresponds to
the main period and it is equal to the external driving force period T
ext
. Additionally, It can be
noticed that now the basic structure consists on a four-lobed shape depicted in Fig. 7.78. The
second main period, equal to T
ext
/3, arises as the distance of two consecutive lobes in the RP
graph. Other subharmonic (1,n) Modes show similar patterns. For instance, Fig. 7.79 shows
7.13. Recurrence Plots and Dynamics 173
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0 2 4 6 8 10
R
e
(

)
Time (s)
T
ext
T
ext
/3
Figure 7.77: Recurrence plots of subharmonic (1,3) Mode. Embedding dimension=1, delay=1,
threshold corridors of 0.1
1
1.5
2
2.5
0 0.5 1 1.5 2 2.5 3 3.5
T
i
m
e

(
s
)
Time (s)
T
ext
1.5
2
0.5 1 1.5
T
i
m
e

(
s
)
Time (s)
T
ext
/3
T
ext
/3
1
2
3
4
Figure 7.78: Main distance in the RP graph of subharmonic (1,3) Mode
the 8-lobe basic structure of the subharmonic (1,5) Mode. For this case, the distance between
consecutive lobes is equal to T
ext
/5. As subharmonicity increases, the number of lobes grow.
In general it is found that for (1,n) modes, the number of lobes is equal to 2(n 1). Figure 7.80
shows the times series and RP of the unstable subharmonic (3,9) Mode. As it is shown, the
basic structures have substantially increased in complexity. Now, the basic pattern consists on
a set formed by various simple forms (Fig 7.81). It can be noticed that the basic structure of the
(1,3) Mode is one of the elements of this new more complex pattern. In this regard, the RP of a
complex mode, such as (3,9), can highlight the content of elementary modes.
7.13. Recurrence Plots and Dynamics 174
Figure 7.79: Main distance in the RP graph of subharmonic (1,5) Mode
-0.2
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0 2 4 6 8 10
R
e
(

)
Time (s)
Figure 7.80: Recurrence plot of unstable subharmonic (3.9) Mode
Quasi-periodic modes
Quasi-periodic modes are the next type of bounded orbits that can be found for harmonic forc-
ing regime. Quasi-periodicity arises when two or more competing frequencies are present in
the dynamics. As it was shown in section 7.9, a chain of stability island can coexist for the
same values of parameters. The island chain grows as the odd integers 1, 3, 5, 7, . . .. It is hence
expected that these patterns appear in the recurrence plots. Figures 7.82 and 7.83b)-c) depict
the recurrence plots for quasi-periodic orbits with a 3, 5 and 7 island structure respectively. In
rst place, it is noticed that the RP of the 3-island structure is qualitatively dierent from the 5
7.13. Recurrence Plots and Dynamics 175
Figure 7.81: Basic structures in the recurrence plot of unstable subharmonic (3,9) Mode
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0 2 4 6 8 10 12 14 16
R
e
(

)
Time (s)
Figure 7.82: Recurrence plot of 3-island quasi-periodic mode. Threshold corridor of 0.1
and 7 island modes, see Figs. 7.82 and 7.83. The recurrence of the 3-island structure is larger
and more complex.
However, patters with a recurrence of 4 (and hence with a periodicity of n = 1 + 4/2 = 3) are
present in the 3-island structure. A careful inspection of Fig. 7.82 shows that there is a trace of
the 4-lobe structures corresponding to the (1,3) Mode. Figure 7.84 shows an amplication of
Fig. 7.82. The pattern shown in Fig. 7.84 is a basic structure of the overall RP. In this shape,
four new ovals have been given birth by a former oval forming a 4-lobe type structure. In this
regard, in the 3-island quasi-periodic structure coexist the periodic (1,1) and (1,3) Modes along
7.13. Recurrence Plots and Dynamics 176
-0.2
-0.1
0
0.1
0.2
0 2 4 6 8 10 12 14 16
R
e
(

)
Time (s)
a) Quasi-periodic 5-islands
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0 2 4 6 8 10 12 14 16
R
e
(

)
Time (s)
b) Quasi-periodic 7-islands
Figure 7.83: Recurrence plots of quasi-periodic modes. Threshold corridor of 0.1
with other more-complex modes.
On the other hand, the basic patterns of the 5-island and 7-island RPs are more similar to the
periodic (1,5) and (1,7) basic structures. The basic 5-island structure is a 8-lobe shape, which
corresponds to an n = 1 + 8/2 = 5 periodicity. The basic pattern of the 7-island structure is a
formerly 12-lobe structure where two of its lobes have migrated into two child lobes (lobes 1

and 7

in Fig.7.85). The lobe 7

has been completely destroyed into two new lobes whereas the
lobe 1

is in the transient blow up state. The process of lobe birth is done through the creation
of recurrence inside the core of a lobe (as shown in Fig.7.85). Then this recurrence grows until
the host lobe breaks into two new lobes.
7.13. Recurrence Plots and Dynamics 177
1.8
2
2.2
2.4
2.6
2.4 2.6 2.8 3 3.2 3.4
T
i
m
e

(
s
)
Time (s)
Figure 7.84: Basic pattern of the quasi-periodic 3-island structure. The central oval has been
shifted into four child ovals forming a 4-lobe structure
3.5
4
4.5
5
5.5
6
6.5
2 2.5 3 3.5 4 4.5
T
i
m
e

(
s
)
Time (s)
1
*
2
3
4
5
6
7
*
8
9
10
11
12
Figure 7.85: Basic pattern of the quasi-periodic 7-island structure. Lobes are numerated. The
lobe 1

is about to blow-up into two new lobes whereas the lobe 7

has already been shifted into


two child lobes
Therefore, the characteristics of quasi-periodic behavior are present in the RP graphs. This lobe
blow-up process picture can be used to understand the quasi-periodicity as a route to chaos as
it will be shown in the next sections.
7.13. Recurrence Plots and Dynamics 178
Chaotic response
The last type of bounded response under harmonic forcing is the chaotic motion. In Fig. 7.86
the recurrence map for a chaotic response is depicted along with their corresponding time series
plots. It can be noticed that the basic structures of quasi-periodic modes are now severely
-0.25
-0.2
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0.25
0 2 4 6 8 10 12 14 16 18 20
R
e
(

)
Time (s)
Figure 7.86: Recurrence plot of a chaotic orbit under harmonic loading with parameters: =
0.245, = 0.98, = 10.0, = 14.0 and initial conditions: X1 = 0.075, X2 = 0.0405
destroyed as it is shown in Fig. 7.87. The presence of a bunch of competing frequencies leads
to a process of lobe destruction and re-birth. However, despite the overall randomness, there
3.5
4
4.5
5
5.5
5 5.5 6 6.5 7
T
i
m
e

(
s
)
Time (s)
Figure 7.87: Basic structure of the RP of a typical chaotic orbit under harmonic forcing regime
is an underlying periodicity (as it is usually the case for chaos). This can be seen from the
7.13. Recurrence Plots and Dynamics 179
chessboard type style that is still present in the plot. In fact, the white block cells indicate
periodicity, although in this case, this periodicity is highly contaminated by the other competing
frequencies.
As it will be shown in the following, chaos under random loading regime presents a qualitatively
dierent pattern when it is examined under recurrence analysis. This fact points out that the
chaos under harmonic motion is substantially dierent than chaos under random loading.
Recurrence plots for earthquake loading
Now, the earthquake input samples are numerically integrated and the RPs corresponding to
the solution time series are computes. It is gauged the eect of the external PGA.
Eect of amplitude
Figure 7.88 shows the RP along with the corresponding time series for two responses under
earthquake loading obtained with the same record and parameters but dierent PGA values.
Both are no-collapse series. First, it is noticed that the patterns of random load response have
little similarity with the ones obtained from harmonic forcing. This fact clearly shows, as it has
been already stated [71], that the RM motion under random loading has qualitatively dierent
behavior than the corresponding to harmonic forcing.
Under random loading regime there are more abrupt changes in the system dynamics than for
harmonic forcing. The clusters (big white blocks) indicate the presence of these changes. When
rocking is initiated, recurrence rates severely decrease, as it is shown in Fig. 7.88 a) and b).
There are three main states in Fig. 7.88 a) which are to be considered; no rocking, transition
rocking and nal rest. At the beginning of the load (0.06.2s), the system accumulates in terms
of vibrations the external energy added by the earthquake input, although there is not RM. Once
a threshold value of the accumulated energy is reached, the rocking initiates and lasts during
and interval of approximately 6 seconds. Finally, the system falls into an attractor and ends up
by dissipating its accumulated energy through impacts.
7.13. Recurrence Plots and Dynamics 180
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
0 2 4 6 8 10 12 14
R
e
(

)
(
o
)

a) PGA = 2.7m/s
2
-5
-4
-3
-2
-1
0
1
2
3
0 2 4 6 8 10 12 14
R
e
(

)
(
o
)

b) PGA = 4.0m/s
2
Figure 7.88: RP for random loading under two amplication factors: a) 1.0 and b) 1.5. The
parameters used were: = 0.245, p = 3.6, = 0.925. An embedding dimension of 1 and a
threshold corridor of 1.0 were used.
However, if the PGA is increased, an additional state appears. In Fig.7.88 b), this state lies
in the interval (11.9 14s), which corresponds to the nal rocking tail observed in the time
series. This nal rocking regime is likely to lead the system to collapse since it has a reservoir
of energy accumulated from the history of the dynamics prior to that state.
7.13. Recurrence Plots and Dynamics 181
Through similar computations using a set of sample functions and dierent PGA values, it has
been notice that, in general terms, the motion under earthquake loading can be divided into the
following parts:
1. No RM.
2. Transition RM.
3. Transition no RM.
4. Unstable RM.
The nal unstable RM state leads to collapse if the load factor is increased from 1.5 to 2.0.
Figure 7.89 depicts the RP of a typical collapse orbit. In this case, the nal unstable state
following the intermediate no RM transition state (interval 7 9s) leads to collapse.
-10
-5
0
5
10
0 2 4 6 8 10 12 14
R
e
(

)
(
o
)

Figure 7.89: RP for random loading of earthquake 20 with PGA=4.0m/s


2
. The parameters used
were: = 0.245, p = 3.6, = 0.925. An embedding dimension of 1 and threshold corridors of
1.0 were used.
Recurrence Quantication Analysis
Quantication of the plots can be performed via the Recurrence Quantied Analysis (RQA)
introduced by Zbilut & Webber [72], which allows to study the intrinsic complexity of dynamical
systems with a few parameters. Those parameters are based on simple pattern recognition
algorithms. Among the various patterns that can be recognized in the RP, the points forming
parallel lines to the main 45
o
diagonal are of special relevance.
7.13. Recurrence Plots and Dynamics 182
RQA is implemented through the algorithms given in [73]. The statistical indicators provided
by the analysis are:
RR Recurrence. This estimator simply counts the black dots in the RP. It is a measure of the
global recurrence.
DET The Determinism parameter is dened as the ratio of recurrence points that form diagonal
structures to all recurrence points. It gives a hint about the determinism of the system.
L The average diagonal length parameter is the average time that two segments of the orbit are
close to each other. It can be interpreted as the mean prediction time [65].
DIV The Divergence estimator is dened as the inverse of the maximum length of diagonal
structures. It is related to the maximum Lyapunov exponent.
LAM The Laminarity is computed as the ratio between the recurrence points forming the ver-
tical structures and the entire set of recurrence points.
TT The Trapping time parameter is dened as the average length of vertical structures. It
measures the mean time that the system will be trapped in a given state.
L
entr
Entropy of diagonal line length distribution.
V
entr
Entropy of vertical line length distribution.
Here, a brief exposition of the RQA parameters has been given which is adequate to the purpose
of the present study. For both denitions and a complete description of these parameters, the
reader is referred to [65] and [73]. In the following, a numerical study is carried out to compute
the parameter values under harmonic and random forced loading.
Harmonic loading
Table 7.6 resumes the main RQA estimators for the periodic modes found by numerical integra-
tion of sine-type input load with constant amplitude. Additionally, Table 7.7 resumes the main
RQA estimators for the unstable periodic modes found by numerical integration. The DIV val-
ues corresponding to both stable and unstable modes are all equal to 5.050 10
4
.
Table 7.8 resumes the main RQA estimators for the periodic modes found by numerical inte-
gration. The DIV values are all equal to 3.127 10
4
.
7.13. Recurrence Plots and Dynamics 183
Mode RR DET DET/RR L TT L
entr
V
entr
(1,1) 0.103 0.097 0.929 106.697 34.545 1.943 1.947
(1,3) 0.147 0.066 0.447 104.941 31.942 1.842 2.638
(1,5) 0.133 0.072 0.542 112.561 7.719 1.596 1.909
(1,7) 0.094 0.104 1.108 100.954 38.811 2.114 0.969
Average 0.119 0.084 0.756 106.288 28.254 1.874 1.866
Table 7.6: RQA estimators for harmonic forcing. Periodic modes
Mode RR DET DET/RR L TT L
entr
V
entr
(1,5) 0.134 0.073 0.549 105.262 35.521 1.989 2.305
(1,7)-I 0.081 0.122 1.505 100.449 38.682 2.120 0.909
(1,7)-II 0.067 0.145 2.156 99.090 37.434 2.160 1.071
(1,8) 0.072 0.137 1.915 98.614 35.755 2.162 1.337
(3,9) 0.156 0.062 0.402 102.974 35.341 2.063 2.308
Average 0.102 0.108 1.305 101.277 36.546 2.099 1.586
Table 7.7: RQA estimators for harmonic forcing. Undamped periodic modes
On the other hand, the values for the RQAestimators corresponding to a chaotic orbit (Fig. 7.86)
are: RR = 0.112, DET = 0.042, DET/RR = 0.375, L = 103.824, DIV 10
4
= 2.501,
TT = 32.559, L
entr
= 1.753, V
entr
= 2.547.
Several trends are noticed when the RQA estimators are compared for the dierent types of or-
bits. The averaged values from the RP estimators are compared in Table 7.9 for the three main
types of bounded orbits in the harmonic forcing regime; periodic, quasi-periodic and chaotic.
In general, the estimator RR is greater for periodic orbits than for undamped, less stable modes.
However, quasi-periodic states show a slightly larger value of RR than periodic modes. Finally,
chaotic orbits exhibit similar RR values (0.113) than that of unbounded periodic modes (0.102).
The DET parameter decreases signicantly from periodic to quasi-periodic and chaotic modes,
although the dierence between these last ones is small. It has been shown in section 7.10 that
7.13. Recurrence Plots and Dynamics 184
Structure RR DET DET/RR L TT L
entr
V
entr
3-island 0.221 0.027 0.123 93.697 23.784 1.488 2.757
5-island 0.124 0.048 0.390 94.103 34.158 1.537 2.395
7-island 0.086 0.071 0.826 94.981 35.916 1.514 1.845
Average 0.144 0.049 0.446 94.261 31.286 1.513 2.332
Table 7.8: RQA estimators for harmonic forcing. Quasi-periodic modes
the quasi-periodicity can be interpreted as a transition to chaos, from periodic stable motion.
When the DET estimator is normalized with the total recurrence points, namely, through the
DET/RR estimator, this trend is more evident. In this regard, DET/RR gives a reasonable hint
about how unpredictable, and hence, chaotic, the system is. The divergence (DIV) parameter
decreases as the motion becomes more unstable. The clear trend observed in Table 7.9 conrms
this fact. The values of L, TT, L
entr
and V
entr
do not exhibit particular trends in relation to the
orbital type. Although there are trends observed for harmonic motion depending on the orbital
Orbit type RR DET DET/RR L DIV 10
4
TT L
entr
V
entr
Periodic 0.110 0.096 1.031 103.782 5.050 32.422 1.986 1.726
Quasi-periodic 0.143 0.049 0.446 94.261 3.127 31.286 1.513 2.332
Chaotic 0.113 0.042 0.375 103.824 2.501 32.559 1.753 2.547
Table 7.9: Comparison of the RP estimators for the three main types of bounded orbits in the
harmonic forcing regime
type, the dierences in values are not large enough to establish general rules. As it will be
shown, the situation for random loading is dierent.
Random loading
The next step consists on the quantication of the RP obtained from numerical integration of
orbits with sample earthquakes. For this issue, the records were numerically integrated and the
7.13. Recurrence Plots and Dynamics 185
corresponding RP was computed. Then, the RQA algorithms were used to measure the estima-
tor values from the RP.
Since, the PGA value is the most important parameter of the external action considered in this
work, the RQA estimators are monitored as the load factor increases in steps of 0.1.
It was noticed that the reliability of the estimate is signicantly improved if the measures were
averaged over the ensemble of samples. Hence, the results correspond to an ensemble average
over the set of samples. The reference parameters = 0.245, p = 3.6 and = 0.925 were cho-
sen for the computations. Figure 7.90 depicts the ensemble mean of the RR estimator against
0.002
0.004
0.006
0.008
0.01
0.012
0.014
2 4 6 8 10 12 14 16
R
R
PGA (m/s
2
)
NO
RM
Figure 7.90: Mean RR versus external PGA in the recurrence plot. West acceleration is indi-
cated by a vertical straight line
the external PGA. As stated above, the RR estimator is directly related to the number of recur-
rences in the RP.
It is clear that once the threshold west acceleration is crossed, the recurrence decreases contin-
uously with PGA. This is due to the fact that for large external load values, the system is less
likely to visit the same states at dierent times.
On the other hand, Fig. 7.91 depicts the ensemble mean of the DET estimator against the ex-
ternal PGA. As it was expected, the determinism of the dynamics decrease when the external
action is magnied. From this and equivalent graphs it can be estimated how chaotic a given
rocking system is under a given earthquake history. For this estimation, the state at the PGA
7.13. Recurrence Plots and Dynamics 186
0.86
0.88
0.9
0.92
0.94
0.96
0.98
1
2 4 6 8 10 12 14 16
D
E
T
PGA (m/s
2
)
NO
RM
Figure 7.91: Mean DET versus external PGA in the recurrence plot. West acceleration is indi-
cated by a vertical straight line
value is referred to the completely deterministic no RM state with DET = 1.0.
Figure 7.92 depicts the ensemble mean of the L estimator against the external PGA. As com-
1
10
100
1000
2 4 6 8 10 12 14 16
L
PGA (m/s
2
)
NO
RM
Figure 7.92: Mean L versus external PGA in the recurrence plot. West acceleration is indicated
by a vertical straight line
mented above, the L can be interpreted as the mean prediction time. In Fig. 7.92 it is noticed
that this deterministic time falls abruptly as soon as the motion is initiated. Then its value de-
creases smoothly with PGA.
7.13. Recurrence Plots and Dynamics 187
The ensemble mean of the divergence estimator as function of the external PGA is depicted in
Figure 7.93. There is a clear increase of DIV with PGA, as observed in the gure. This result
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
2 4 6 8 10 12 14 16
D
I
V
PGA (m/s
2
)
NO
RM
Figure 7.93: Mean DIV versus external PGA in the recurrence plot. West acceleration is indi-
cated by a vertical straight line
was expected; the divergence of nearby trajectories increases with the instability and, hence,
with PGA.
Laminarity (LAM) and trapping time (TT) are of special importance for the present study since
these parameters are able to estimate the number of unstable periodic modes, which are directly
related to the stability of the problem. Figure 7.94 depicts the ensemble mean of the LAM es-
timator against the external PGA. It is noticed that the presence of laminar states continuously
decrease with PGA. Then the system is less likely to remain in a given state for an interval of
time.
On the other hand, Fig. 7.95 depicts the ensemble mean of the TT estimator against the external
PGA. There is a fast decrease of the trapping times as the external load increases. This param-
eter is related with the mean rst passage time introduced in section 7.5. In fact, if Fig. 7.95 is
compared against Fig. 7.30 a strong correlation can be observed between both graphs.
Figure 7.96 depicts the ensemble means of the L
entr
and V
entr
estimators against the external
PGA. These parameters, related to the longitudinal and vertical entropies are more dicult to
interpret. In Fig. 7.96 a) it is noticed that once the threshold value of the west acceleration
7.13. Recurrence Plots and Dynamics 188
0.92
0.93
0.94
0.95
0.96
0.97
0.98
0.99
1
1.01
2 4 6 8 10 12 14 16
L
A
M
PGA (m/s
2
)
NO
RM
Figure 7.94: Mean LAM versus external PGA in the recurrence plot. West acceleration is indi-
cated by a vertical straight line
4
6
8
10
12
14
16
18
2 4 6 8 10 12 14 16
T
T
PGA (m/s
2
)
NO
RM
Figure 7.95: Mean TT versus external PGA in the recurrence plot. West acceleration is indi-
cated by a vertical straight line
has been crossed, the value of L
entr
blows-up to a value of approximately 2.4. Then it remains
almost constant for a range of PGA values (a
w
8.0m/s
2
) approx. Finally, after this almost
constant state it decreases continuously with PGA. On the other hand, the entropy associated to
the vertical structures (Fig. 7.96 b)), decays smoothly with PGA once the rst jump has been
achieved.
7.14. Summary 189
2
2.1
2.2
2.3
2.4
2.5
2 4 6 8 10 12 14 16
L
e
n
t
r
PGA (m/s
2
)
NO
RM
a) L
entr
0
0.5
1
1.5
2
2.5
2 4 6 8 10 12 14 16
V
e
n
t
r
PGA (m/s
2
)
NO
RM
b) V
entr
Figure 7.96: Ensemble means of the L
entr
and V
entr
estimators versus external PGA in the recur-
rence plot. West acceleration is indicated by a straight line
7.14 Summary
This section summarizes the main results obtained from the dierent approaches contained in
this chapter.
Maximum angles and overturn
From the computations of the correlation maxima it can be concluded that there exist a corre-
lation between consecutive maxima for harmonic forcing. This correlation is maintained for
random loading.
On the other hand, correlation decays fast as the number of impacts between maxima increase.
This fact suggests that if a reasonable model for the probability of maxima is to be addressed,
consecutive maxima should not be considered as independent random variables.
By considering the ensemble means of the angle and velocity it has been found that there are
intervals of time for which these quantities remain almost stationary. Since the energy is a func-
tion of and

, the same stationarity regime was found for the energy.


On the other hand, from the computations of the energy of the block it is found the for a given
earthquake record, higher PGA values do not necessarily imply larger energies. However, it can
7.14. Summary 190
be veried through numerical simulation that, in the stationary limit of the ensemble mean of the
mechanical energy increases with PGA. Furthermore, it has been found that energy decreases
exponentially with damping for a given value of the external PGA.
Impacts and stability
The analysis of the number of impacts for dierent values of the system parameters and the
external PGA have shown non trivial relationships between the number of impacts per time
unit N

and the parameters. Through numerical computation the following observations can be
made:
1. N

decreases as a power of (PGA-a


w
).
2. N

increases as a power of .
3. N

depends strongly on p in a nonlinear fashion.


4. N

does not change signicantly with .


On the other hand, it has been stated that N

reaches a stationary limit n


p
at which its value is
almost independent of time. An expression for the dependence of n
p
with the external PGA has
been found.
First Passage Time
A functional expression has been found for the mean exit time < FPT >. This quantity may be
used as a rst estimator of the mean collapse time. However, the time at which the angle equals
(FPT) is not the time at which the structure collapses.
Stability in the amplitude-frequency range
The non-overturning orbital types are classied into three main types: periodic, quasi-periodic
and chaotic, according to [47].
An extensive numerical computation has been carried out for harmonic forcing regime in the
7.14. Summary 191
amplitude-frequency space for dierent values of the system parameters. Then, stability bound-
aries in the range where derived. The main conclusions from that study are:
The number of collapsed orbits increases fast with .
The collapse rates increase with the frequency parameter p. This implies that if two
blocks with dierent size but equal slenderness ratio are subjected to the same external
action, the larger block may survive whereas the smaller topples. This is the so-called
scale eect rst discovered by Housner [7].
Damping strongly modies the instability boundaries in the map; high damping
considerably reduces the collapse rates.
The external phase of the input signal was shown to have important eects in the
stability boundaries. This is due to the fact that physically, a cosine-type signal is not
equivalent to a sine-type signal.
On the other hand, the stationary parts of the response obtained with a cosine-type and
sine-type functions should be identical. This fact points out that the transient regime of
the response is the most relevant part of the orbit concerning collapse.
Bifurcation thresholds
Through numerical computations threshold values in the system parameters have been found,
for which the system characteristics change abruptly. The control parameter chosen in order to
monitor the changes in the dynamics is the period of the response, computed through a Fourier
transform analysis.
Nevertheless, not all the orbits were periodic and, hence, the computed periods by means of
the Fourier transform of the response can only give an estimate to the period. This indicates a
strong limitation of the period analysis.
A sequence of period jumps was found by ranging the system external amplitude and frequency:
1-3,3-4,4-5,5-6.5. However this sequence does not entirely reproduce the odd integer sequence
of subharmonic (m, n) modes found through numerical computation. This is due to the fact
that for non-periodic orbits (i.e. quasi-periodic and chaotic), the periods are not well dened
7.14. Summary 192
quantities. Fourier transforms then give an average of the orbital periods.
Other sequences in the period jumps have been found when the system parameters are varied.
However no explanation could be found for these jumps.
Delay reconstruction analysis
Through the delay-coordinates the main orbital types were analyzed for various delay lags. It
was found that the subharmonic modes could only exist for denite delay intervals.
Poincar Surface of Section analysis
Poincar Surface of Section analysis allowed to identify the structure of the phase space. The
main orbital types were related to their corresponding PSS: Periodic orbits consist on a nite
number of Poincar points, quasi-periodic modes result in the so called stability islands (or
simply islands) in the PSS. Finally, chaotic behavior traces a fuzzy-noisy pattern in the PSS.
The PSS were computed for harmonic forcing with constant amplitude and for random loading.
Additionally, the eect of white noise on the orbits and the PSS was examined. For harmonic
forcing regime, the following results were obtained:
1. The intersections between the stable and unstable manifolds lead to accumulation of
points in wave patterns. These intersections occur in the domains: ( > 0, r > 0) and
< 0, r > 0, when the absolute value of the rocking angle increases. The intersection
patterns are severely aected by the amplitude and frequency.
2. It was found that, for harmonic loading, dierent sequences of attractors and stability
islands coexist for the same values of the system parameters. These attractors are due to
quasi-periodic behavior and their sequences follow the odd integer sequence: 1, 3, 5, 7,
etc.
3. When damping is present, the stability islands are either destroyed or shifted into at-
tractors. With no damping at hand, no periodic structure is observed; only chaotic and
quasiperiodic patterns are found. It was found that at every impact there is a reduction of
the phase space volume by an amount of
2
, equal to the energy reduction.
7.14. Summary 193
4. For damped systems, odd n values of subharmonic (1, n) modes coexist at dierent exter-
nal amplitude levels.
5. When no damping is allowed and for small values of the external frequency (typically
< 6), a single attractor with fractal structure is observed at the origin.
6. It is noticed that as decreases, more quasi-periodic modes coexist. For undamped mo-
tion, it is noticed that Quasiperiodicty only appears for certain values of .
7. When the external phase is varied, the structure of the phase space persists.
When white noise was added to the harmonic input signals, the phase space structure was par-
tially destroyed, although some of the underlying scheme persisted. Along the numerical com-
putation it was observed that the noise does not aect all the orbits in the same proportion.
Periodic modes were more robust to noise than quasi-periodic and chaotic responses.
It was observed that if high levels of noise are allowed, the phase space patterns are fully de-
stroyed in the associated PSS. However, the single attractor at the origin persisted even for high
values of noise. Under random loading, the structure of the phase space was severely altered.
The main conclusions derived form the study of PSS under random loading regime are:
1. If no damping is allowed, the Poincar points are equally distributed within the stable
region determined by the separatrix. As damping increases, the points collapse to the
attractor at (0,0).
2. As the external amplitude increases, the Poincar points scatter over the (X1, X2) plane.
3. The attractor structure for random loading regime consists on three main regions:
1) Dense core, 2) Scattered region and 3) Diagonal wings.
4. It has been found that the probability of falling within the stable domain decreases rapidly
with .
5. The attractor becomes more narrow and large in the vertical direction as p increases due
to the phase-space contraction. Additionally, the attractor rotates counter-clockwise as p
increases.
The separatrix crossing rate was computed for dierent PGA values. It was noticed that the
so-called diusion mechanism increases smoothly until PGA a
w
. Once the West acceleration
7.14. Summary 194
is reached, the crossing rates increase fast.
Finally, a Gaussian probability was used to investigate the distribution of the Poincar points
in the attractor. Although the model was very simple and could not give account for the wing
eect it suced to estimate the basic distribution in the attractor.
Recurrence Analysis
Recurrence analysis allowed to identify patterns and periodicity for the dierent types of or-
bits under harmonic forcing regime. On the other hand, the quantication of RPs for random
loading showed to have denite trends with the external PGA. This fact allowed to identify the
"DET" estimator as a suitable indicator for the quantication of chaos in the rocking block dy-
namics for a given value of PGA.
For harmonic forcing the main results observed are:
1. The stationarity of both periodic and quasi-periodic motion is reected in the homogene-
ity of the plots. The overall checkerboard aspect indicates the inherent periodicity of the
system.
2. Quasi-periodic behavior exhibits similar patterns to periodic motion, although more intri-
cate. The characteristics of quasi-periodic behavior are present in the RP graphs through
a lobe destruction and oval birth process. This schematic picture can be used to under-
stand the quasi-periodicity of the system. Furthermore, since chaotic orbits were found
to have a high rate of lobe-destruction process, quasi-periodic motion can be understood
as a transition from periodic to chaotic motions.
The following conclusions are derived for the case of random loading:
1. The RPs of randomloading change substantially with respect to harmonic forcing regime.
Laminar (states which do not change or change slowly) states occur frequently as it is indi-
cated in the RPs by the presence of horizontal and vertical lines. These states are missing
in the harmonic forcing regime. On the other hand, abrupt changes in the dynamics are
also evidenced by the presence of big white blocks (clusters) in the RP.
7.14. Summary 195
2. The RPs allowed to identify three main parts for an orbit resulting from random motion
regime: First, there is a no-rocking range, then there is a transient rocking following for
a transient no-RM state. After this, there is an intense RM which may lead to overturn.
The abrupt changes in the system dynamics (clusters in the RP) are interpreted as an
indicator of chaos.
3. There are clear patterns found in the RQA measures. In this regard recurrence quanti-
cation analysis has shown to be an adequate tool for the investigation of RM dynamics
under random loading. However, it has been noticed that for two dierent input signals
(accelerograms) with the same PGA, the RQA estimators may give contradictory results.
On the other hand, if the estimators are averaged over the ensemble of input series, the
results are reasonable. Therefore, an ensemble of orbits (and, hence RPs) must be used
in order to interpret correctly the RQA measures.
4. Among all RQA estimators, the determinism "DET" has found to be the best indicator of
chaos for RM dynamics since it decreases monotonically, almost linearly, with PGA once
the West acceleration is reached.
Part IV
Applications to Earthquake Engineering
196
Chapter 8
Probability of Collapse for SDOF
Structures
8.1 Introduction: The Rocking Spectrum
Following a standard approach [39],[41], the response of the system X() is understood as a
stochastic process given by the following stochastic dierential equation:

X() = F[a
g
()] (8.1)
In this expression, F represents, in general, a nonlinear lter and a
g
is regarded as the normal-
ized input acceleration. The input signal a
g
is itself a stochastic process whose main character-
istics are directly related to the system response [74].
The RM spectrum
Earthquake response spectrum plays an essential role in everydays earthquake engineering for
both design criteria and earthquake risk assessment. The design spectrum plots the maximum
amplitude of a linear oscillator under an external action as function of the inner frequency and
for a xed value of damping (typically 5%). Most regulation codes include certain behavior
factors in order to account for the divergences existing between linear elastic models and real
197
8.1. Introduction: The Rocking Spectrum 198
structures. However, if these design spectrum plots are to be used for masonry structures, a
careful analysis should be taken into consideration when rocking is present.
Makris & Konstantinidis [8] highlighted the dierences between the oscillatory response of a
SDOF oscillator and the rocking response of a slender rigid block. They showed that for rocking
structures, the simple earthquake design approach should be abandoned, since it led to false
conclusions, in particular for smaller, less-slender blocks.
Figure 8.1 a) shows the rocking spectrum corresponding to two values of . For that plot, the
|
max
| were computed as function of 1/p. On the other hand, the equivalent-linear-oscillator
0
2
4
6
8
10
12
14
0 1 2 3 4 5 6
<
|

|
m
a
x
>
2/p (s)
Rocking Spectrum: =0.245
=1.0
=0.925
a) Rocking spectrum b) Equivalent linear oscillator response spectrum
Figure 8.1: The rocking spectrum compared with the equivalent linear oscillator response spec-
trum of synthetic earthquakes used along this work
response spectrum of the synthetic earthquakes used along this work is shown in Fig. 8.1 b).
From Figs. 8.1 a) and 8.1 b) it can be noticed that the frequency parameter p plays a equivalent
role to the period in the typical response spectrum graph. However, it can be noticed that
the "platform" of the linear-equivalent-oscillator spectrum (Fig 8.1 b)) is missing in the RM
spectrum. That constant interval is normally used for design purposes. The presence of peaks
in the RM spectrum along with the nonlinearity of RM motion conrms the conclusion of [8].
For these reasons, a probabilistic approach to the collapse of rocking blocks under earthquake
loading is the more adequate way to tackle the assessment of block structures under earthquake
loading.
8.2. Experimental Collapse 199
8.2 Experimental Collapse
A set of experimental tests on specimens 3 and 4 was conducted in order to estimate the em-
pirical probability of collapse. The tests were carried out with increasing load factor, until
collapse was reached. Table 8.1 shows the number of collapses observed for specimen 3. From
PGA(m/s
2
) Num. Experiments Num. Collapses Probability
1.29 8 0 0
1.72 18 2 0.110
2.15 28 6 0.214
2.58 26 11 0.423
3.02 18 8 0.444
3.45 11 8 0.727
Table 8.1: Number of observed overturns for specimen 3.
this values, it is clear that the ratio between number collapses and number of tests (that is, the
probability of collapse) increases with the external acceleration. The same trend was observed
for specimen 4. In section 8.3 these results will be compared with the probability of collapse
computed from numerical simulation. Then, it will be possible to achieve adequate criteria for
the safety of a rigid block under earthquake actions.
Characteristics of collapse orbits
Figure 8.2 a) shows the absolute value of the measured angle of specimen 3 for a typical collapse
orbit. In the graph, the theoretical and experimental values of are also indicated. From this
and other similar tests it has been observed that, before the collapse occurs, there is a clear
instability band from which no return to bounded motion is possible. From the conclusions
of section 7.14 this lag is interpreted as the oscillations of the perturbed unstable manifold W
u
around the stable manifold W
s
. This threshold lies close to the static maximum angle , which
corresponds to the separatrix introduced in section 7.10.
8.2. Experimental Collapse 200
0
5
10
15
20
6 7 8 9 10 11 12
|

|
(
o
)
(s)
teor
fit
Collapse orbit for specimen 3
-10
-5
0
5
10
15
20
25
14 14.5 15 15.5 16 16.5
(s)

(o)
vel/10
acc/100
b) Elements of a typical collapse orbit
Figure 8.2: Elements of a typical collapse orbit; angle, velocity and acceleration.
Figure 8.2 b) depicts the angle (absolute value), velocity and acceleration of a typical collapse
orbit
1
. For this case, the selected record was earthquake 1 with a PGA=4.5m/s
2
on specimen 4.
As stressed above, once the instability boundary is reached, the angle starts to diverge. However,
a fact must be also highlighted concerning the velocity; within the instability region, its value
remains close to zero. This fact can also be veried by comparing the phase plots of a collapsed
and a non-collapsed orbit.
Figure 8.3 depicts the phase plots for specimen 4 under earthquake 1 with a 4.5m/s
2
of PGA
value. The accumulation of points close to the horizontal axis corresponds to a velocity value
which remains close to zero. Similar results are obtained for the rests of collapsed orbits.
Finally, Fig. 8.4 depicts the mechanical energy against the absolute value of the rocking angle
in order to highlight the strong correlation between these magnitudes. It can be noticed that the
maximum of energy is proportional to the maximum of ||.
Regarding the relationship between the static angle and the maximum angle for the non-
collapsed orbits it was observed that specimens 2 and 3 have |
max
| lower than . However,
for specimen 4, the maximum value obtained from all the experimental maximum angles was
21.14
o
, which is larger than the theoretical (17.7
o
) and tted (15.35
o
) values of . These results
1
Velocity and acceleration include a factor of 1/10 and 1/100 respectively in order to t the width limits of the
orbits
8.2. Experimental Collapse 201
-100
-50
0
50
100
-10 -5 0 5 10
d

/
d
t
(o)
a) No collapse
-100
-50
0
50
100
-10 -5 0 5 10
d

/
d
t
(o)
b) Collapse
Figure 8.3: Comparison of phase orbits for collapsed and no collapsed orbits: a) No collapse. b)
Collapse. The accumulation of points close to the horizontal axis in b) indicates that the value
of the velocity remains close to zero
0
0.01
0.02
0.03
0.04
0.05
0.06
4 6 8 10 12 14 16
E
n
e
r
g
y
(s)
||/10
Energy
Figure 8.4: Mechanical energy of a typical collapse orbit
point out that the dynamic collapse threshold is lower than the static. This will be discussed in
detail in the following section.
8.3. Probability Curves for Overturn 202
8.3 Probability Curves for Overturn
The aim of this chapter is to establish the relationship between the probability of collapse, and
the peak ground acceleration (PGA) and the system parameters (, p, ).
Probability of collapse. Numerical study
The numerical probability of collapse P
c
is computed using the following procedure:
1. For a given set of parameters, the ensemble of samples is numerically integrated. The
following set is chosen as reference values:
= 0.245 ( = 4),
p = 3.6,
= 0.925.
Then, two of these parameters are xed while the remaining is varied.
2. The absolute value of maximum angles is then computed for each output series.
3. The number of series whose maximum angles exceed /2 is stored.
4. The ratio of collapsed orbits is regarded as the numerical probability of collapse.
Eect of the slenderness ratio
The eect of the slenderness ratio on the probability of collapse P
c
is numerically investigated
through the procedure described above. Figure 8.5 a) shows the values of P
c
for p = 3.6 and
= 0.925 and dierent PGA values. Figure 8.5 b) shows the P
c
curves for xed PGA values. It
can be noticed that for a given PGA, the probability of collapse increases with the slenderness
ratio.
Eect of the frequency parameter
The dependence of P
c
with p is investigated for = 0.245 and = 0.925 and shown in
Fig. 8.6. From these plots it can be noticed that for a given PGA value, the probability of
collapse increases with p.
8.3. Probability Curves for Overturn 203
0
0.2
0.4
0.6
0.8
1
0 2 4 6 8 10
P
r
o
b
a
b
i
l
i
t
y

o
f

C
o
l
l
a
p
s
e
PGA (m/s
2
)
=3
=6
=9
a) P
c
for dierent PGA values.
0
0.2
0.4
0.6
0.8
1
0 1 2 3 4 5 6 7 8 9
P
r
o
b
a
b
i
l
i
t
y

o
f

c
o
l
l
a
p
s
e

PGA=4.05m/s
2
PGA=4.32m/s
2
PGA=5.40m/s
2
PGA=6.21m/s
2
PGA=7.29m/s
2
b) P
c
for xed PGA
Figure 8.5: Dependence of the probability of collapse with the slenderness ratio
0
0.2
0.4
0.6
0.8
1
0 2 4 6 8 10
P
r
o
b
a
b
i
l
i
t
y

o
f

C
o
l
l
a
p
s
e
PGA (m/s
2
)
NO RM
p=3.0
p=3.6
p=5.0
0
0.2
0.4
0.6
0.8
1
3 4 5 6 7 8 9
P
r
o
b
a
b
i
l
i
t
y

o
f

c
o
l
l
a
p
s
e
p(Hz)
PGA=3.51m/s
2
PGA=4.05m/s
2
PGA=4.32m/s
2
Figure 8.6: Dependence of the probability of collapse with the frequency parameter
Eect of damping
Damping eects on collapse are investigated for = 0.245 and p = 3.6 and shown in Fig. 8.7.
For a xed PGA value, it can be seen that P
c
increases with .
-curves for P
c
Through algebraic manipulation, all the parameters can be expressed in terms of the slenderness
ratio :
8.3. Probability Curves for Overturn 204
0
0.2
0.4
0.6
0.8
1
0 2 4 6 8 10
P
r
o
b
a
b
i
l
i
t
y

o
f

C
o
l
l
a
p
s
e
PGA (m/s
2
)
NO RM
=0.80
=0.90
=0.99
0
0.2
0.4
0.6
0.8
1
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95
P
r
o
b
a
b
i
l
i
t
y

o
f

c
o
l
l
a
p
s
e

PGA=4.05m/s
2
PGA=4.32m/s
2
PGA=5.40m/s
2
Figure 8.7: Dependence of the probability of collapse with the damping coecient
= tan
1
(
1

),
p =
1
2
_
3gcos(),
= 1
3
2
sin()
2
.
With these values, curves for the probability of collapse as function of only are integrated for
dierent PGA values. Figure 8.8 shows the results obtained from those computations.
0
0.2
0.4
0.6
0.8
1
0 2 4 6 8 10
P
r
o
b
a
b
i
l
i
t
y

o
f

c
o
l
l
a
p
s
e
PGA (m/s
2
)
=4 =5 =6
=7
=8 =9 =10
Figure 8.8: Probability of collapse as function of
8.4. Analytical Model for the Probability of Collapse 205
Numerical and experimental probability of collapse. Comparison
In order to build a numerical model for the probability of collapse under dierent conditions,
validation of the model with empirical tests is examined. The experimental probability of col-
lapse for specimen 3 was reported in table 8.1. Now, Figs. 8.9 a) and 8.9 b) show the comparison
between experiment and numerical simulation of the overturn probability for specimens 3 and
4 respectively. Very good agreement is found.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 1 2 3 4 5 6
P
r
o
b
a
b
i
l
i
t
y

o
f

C
o
l
l
a
p
s
e
PGA(m/s
2
)
a
w
NO RM
Numerical
Experiment
a) Specimen 3
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 1 2 3 4 5 6 7 8
P
r
o
b
a
b
i
l
i
t
y

o
f

C
o
l
l
a
p
s
e
PGA(m/s
2
)
a
w
NO RM
Numerical
Experiment
b) Specimen 4
Figure 8.9: Comparison between numerical and experimental probability of collapse for speci-
men 3 and 4.
8.4 Analytical Model for the Probability of Collapse
Introduction. Collapse criteria
Through the preceding chapters it has been shown how the problem of stability of rocking struc-
tures can be extremely sensitive to variations in the system parameters and boundary conditions.
In addition to this fact, knowledge of a real structure parameters is often scarce or impossible.
Therefore, the overturn of masonry structures can only be addressed from a probabilistic point
of view.
8.4. Analytical Model for the Probability of Collapse 206
Dierent approaches have been developed in order to tackle the problem of collapse of rigid
blocks under external loading. The main approaches, analyzed next, are:
Direct energy balance criteria,
Transition to chaos as an indicator of instability, and, hence, collapse.
Stochastic analysis of the associated diusion equation.
The probability of collapse is determined by numerical integration of the equations of
motion.
Energetic balance
In order to understand collapse it is necessary to revisit to Eq. 7.12: E = K+V
e
(). The collapse
is understood as the probability of exceedance of the total energy E() of the maximum of the
time-dependent potential V
e
(). This maximum depends on time.
From the kinetic energy theorem it is found that the increment of kinetic energy is equal to the
total work done to the system:
K = W (8.2)
If this work W is divided into those done by conservative forces (i.e. those which are derivable
from a potential) and non-conservative forces (in this case, impacts);
W = W
c
+ W
nc
(8.3)
where W
c
represents the work done by the conservative forces while W
nc
is the work done by
impacts. The conservative work is the sum of the work done by gravitational forces W
g
and
the work of the earthquake action W
e
. Additionally, W
g
= U, where U is the gravitational
potential (Eq. 5.16). Therefore, the increment of kinetic energy between the rest and collapse
positions holds:
K = U + W
e
+ W
nc
(8.4)
8.4. Analytical Model for the Probability of Collapse 207
In order to derive conditions for collapse, Housner [7] made the following assumptions:
1. The velocity of the block is zero when it topples.
2. There are no impacts within collapse interval.
3. The blocks are slender.
4. The displacements are small.
With these assumptions, Eq. 8.4 reduces to:
W
e
= U (8.5)
and Housner [7] stated that collapse occurs when:
W
e
U (8.6)
From Eqs. 5.40 and 5.16 of it is found that: Q
e
r
= (MgR)a
g
cos() cos(r) and U = MgR(1
cos ). By dening the maximum of the nondimensional eective potential e
c
as:
e
c
1 cos() (8.7)
and, substituting these quantities in Eq. 8.6, the following holds:
_

0
a
g
() cos() cos( r)dr e
c
(8.8)
Under the assumption of small displacements and slender blocks, cos(r) 1, and 1cos()

2
/2. Therefore, the preceding equation can be recast as:
_

0
a
g
()r

2
2
(8.9)
where

is the time of collapse, equal to the exit time dened in section 7.5.
As it was shown in chapter 2, if the external action is constant or sinusoidal, the equations
of motion for slender blocks can be solved for small angles (Eq. 2.10). With this procedure
Housner was able to determine the maximum amplitude of the excitation and

[7]. The most


unsatisfactory hypothesis is that the block has a zero angular velocity when the static collapse
angle is reached [9].
8.4. Analytical Model for the Probability of Collapse 208
On the other hand, if the external action is considered as a random force and impacts are in-
cluded into the model, no analytical solution has still been provided for the probability of col-
lapse. In this case, collapse should be understood as a typical escape problem. Figure 8.10
depicts the static potential well given by Eq 5.21, where the maximum of the eective potential
is e
c
1 cos(). For free RM, it is clear that collapse is only possible if the angle equals the
0
0.005
0.01
0.015
0.02
0.025
0.03
0.035
0.04
0 0.05 0.1 0.15 0.2 0.25 0.3
v
e
(
r
)
r
Static collapse

E
v4(x)
1-cos()
Figure 8.10: Bounded states in the eective potential. The maximum e
c
= 1cos() is indicated
by a straight line.
critical angle . However, when external forcing is present, the shape of the well is deformed
and "escape" from the well is possible for values lower than .
Stability thresholds crossing
The overturn of rigid blocks has been also interpreted as instability, i.e. chaotic motion. In this
context, the eorts were focused in the determination of the conditions of transition to chaos.
Within the context of the dynamics in the phase plane, the safe domain is the separatrix which
divides the set of orbits into stable and unstable (Figs. 7.44 and 7.66). When a harmonic force
was assumed, analytical tools such as Melnikov method were used to nd the conditions of
8.4. Analytical Model for the Probability of Collapse 209
chaos along with the distance between the stable and unstable manifolds [13],[15]. Recently,
the analysis of collapse fromthe dynamical systems point of view, [67], has also been addressed.
Stochastic Analysis
More recently, attention has been given to the stochastic nature of the problem, by using stochas-
tic analysis, [71],[75],[76],[77].
Numerical integration of P
C
Iyengar and Monahar [78] studied the overturning of the block as a First Passage Time problem
(FPT) using the diusion equation. But most authors [9],[41],[42] dealt directly with the more-
realistic random forces. For this case, the probability of overturn under earthquake actions was
found by numerically integrating the equations of motion.
Proposed models for collapse
The common characteristic of all the above approaches is the exceedance of a given "safe" do-
main. In the present contribution a middle of the road viewpoint is adopted. The main thesis is
the close link between the complexity of the system at hand and the need for a probabilistic ap-
proach. Complexity is regarded as an intrinsic phenomenon generated by the high nonlinearity
of the system. The main consequence of this complexity is the onset of chaotic dynamics.
As it has been shown along this work, the presence of pronounced sensitivity both in parameters
and initial conditions is unavoidable. Moreover, for a given parameter value, a pair of initially
nearby states will diverge if a bifurcation is present (section 7.7). The rate of this divergence is
an intrinsic property of the dynamics, referred as the maximum Lyapunov exponent.
There are two events which are relevant to collapse:
RM Initiation of rocking
C Collapse
Obviously, the event "C" is included in the event "RM"; C RM and hence, C RM = C. The
probability of collapse is the compound event of "RM" and "C": P
C
= P
CRM
. If "P
C|RM
" labels
8.4. Analytical Model for the Probability of Collapse 210
the conditional probability of collapse once the conditions of initiation of rocking are satised:
P
C
= P
C|RM
P
RM
(8.10)
The event "RM" of initiation of rocking is expressed as (Eq. 6.1)

RM
; a
g
(
RM
) a
w
/g (8.11)
That is, there is a time
RM
for which the external acceleration exceeds the west acceleration. It
must be noticed that this event is equivalent to the event:
max(a
g
) a
w
/g (8.12)
Therefore the probability P
RM
can be expressed as:
P
RM
= P[PGA a
w
] (8.13)
In a model which has the forcing amplitude as variable, this probability is a step function:
P
RM
= (PGA a
w
) (8.14)
where (x) is the Heaviside step function as dened in appendix C. The main eort here is
given to nd the probability P
C|RM
, once the external acceleration exceeds a
w
.
In the following two models for P
C|RM
are compared and discussed. One is based on the mean
exit time whereas the other is based on the stationary solution of the associated Fokker-Planck
equation.
Poisson model for P
C|RM
There are two parameters regarding the collapse process which are relevant in a Poisson model;
the probability P
C|RM
(t) of exit from the potential well over a given time interval [0, t] and the
mean exit time < T
e
>.
As it is known [74],[79], Poisson processes consider exits as rare events in their probability
distribution laws. In this regard, collapse is assumed as a unlikely outcome in a sample space
consisting on a set of trials for each interval of time. However, it is both an experimental and
8.4. Analytical Model for the Probability of Collapse 211
numerical evidence that the probability of collapse increases with PGA. Then, there must be a
coupling between the mean exit time < T
e
> and the PGA.
The probability time for the Poisson process is given by
P
C|RM
(t) = 1 e
t/<T
e
>
(8.15)
In section 7.5, a functional expression (Eq. 7.7) was derived for the expected value of the time.
However, as commented at the end of that section, the < FPT > corresponds only to the exit
time from the static eective potential (see Fig. 8.10), as expressed in Eq. 7.6.
On the other hand, closed form solutions of < T
e
> are dicult to obtain since, despite the
eorts [15],[41],[67],[71], collapse criteria of rocking structures under general random loading
is still an active eld of research.
As a rst choice for < T
e
> is to assume < T
e
>=< FPT >. However, this model did not report
an adequate tting when it was tested against numerical simulation. This fact is interpreted as
the error when equating < T
e
> to < FPT >.
A more suitable approach is to consider:
< T
e
>= G(PGA a
w
)
1

PGA a
w
(8.16)
where G(PGA a
w
) is a function of the acceleration dierence only. By assuming a potential
law in G(PGA a
w
), the conditional probability may be expressed as:
P
C|RM
= 1 e
C
1
(PGAa
w
)
C
2
(8.17)
where the constants C
1
and C
2
depend, in general, on (, p, ) and
f
is the duration of the
external loading. Therefore, from Eqs. 8.10 and Eq. 8.14, it is possible to obtain:
P
C
= (PGA a
w
)(1 e
C
1
(PGAa
w
)
C
2
) (8.18)
Figure 8.11 shows the comparison between the modied Poisson model of Eq. 8.18 and numer-
ical simulation with C
1
= 0.132, C
2
= 0.230.
Markovian Model. The Fokker-Planck equation
As it was found in chapter 7, the maximum of the angles are highly correlated variables. There-
fore, a probability model for ||
max
must take this fact into account. As a rst approximation, a
8.4. Analytical Model for the Probability of Collapse 212
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 2 4 6 8 10 12
P
r
o
b
a
b
i
l
i
t
y

o
f

c
o
l
l
a
p
s
e
PGA (m/s
2
)
NO RM
Numerical Simulation
Modified Poisson Model
Figure 8.11: Modied Poisson model for the probability of collapse for C
1
= 0.132, C
2
= 0.230
Markov-chain model is proposed. The derivation of the associated Fokker-Planck equation is
presented in annex D.
Stationary solution. Model for P
C
With the drift and diusion terms derived in annex D the stationary solution of the Fokker-
Planck equation is:
P
C
= (PGA )

1
c
a
_
e
c
0
exp

2ln()
a
e
2

j
(t t
j
)

(8.19)
Although a number of restrictive hypotheses have been made, Eq. 8.19 is not still solvable
in a closed-form solution. Part of the problem stems from the coupling between z and a
g
in
the diusion coecient. Furthermore, both drift and diusion coecients are time-dependent,
which clearly avoids analytical solutions.
8.4. Analytical Model for the Probability of Collapse 213
On the other hand the denite integral of an exponential function of x
2
has no analytical solution.
In section 7.2 it was found that the expected value of the energy reaches a stationary value e
(Eq. 7.1): e = (p, , )(PGA a
w
). As a further step towards the solution, it can be assumed
that,
1. < z
2
>
2
.
2. a(, t) is assumed as constant.
3. The energy in Eq. 8.19 is substituted by its stationary value e found in section 7.2.
4. The expected value of F(t) is substituted by its temporal average:
F(t) =
1
T
_
T
0
F(t

)dt

=
1
T

j
_
T
0
dt

(t

t
j
) (8.20)
where the integral is extended over the duration of the motion. However,
1
T

j
_
T
0
dt

(t

t
j
) =
N(T)
T
(8.21)
where N
T
is equal to the number of impacts per time unit in the interval of non-dimensional
time T. Since T = p, it holds:
F(t) =
N

p
(8.22)
where, N

is the number of impacts per time unit (Eq. 7.5 of section 7.4). By taking the
stationary limit of N

:
F(t) = n
0
+
c
1

PGA a
w
(8.23)
Under these assumptions, Eq. 8.19 renders:
P
C
= (PGA )(1 d
1
exp[c
1
(PGA a
w
)
2
+ c
2
(PGA a
w
)
3/2
]) (8.24)
where "d
1
" , "c
1
" and "c
2
" are constants which may depend on system parameters (, p, ) but
not on external parameters (PGA, S
a
g
()).
Normalization of Eq. 8.24 is implemented by the requirement that the conditional probability
P
C|RM
is zero if PGA = a
w
. Then, it holds that d
1
= 1, and using Eq. 8.24, it is possible to
obtain:
P
C
= (PGA )(1 exp[c
1
(PGA a
w
)
2
+ c
2
(PGA a
w
)
3/2
]) (8.25)
8.4. Analytical Model for the Probability of Collapse 214
Numerical simulations have shown that Eq. 8.25 yields very good results. Figure 8.12 shows
the model for the probability of collapse of Eq. 8.25 with values c
1
= 0.42 and c
2
= 0.12.
The expression, although not entirely exact (due to the restrictive hypothesis assumed), it may
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 2 4 6 8 10 12
P
r
o
b
a
b
i
l
i
t
y

o
f

c
o
l
l
a
p
s
e
PGA (m/s
2
)
NO RM
Numerical Simulation
Fokker-Plank Model
Figure 8.12: Fokker-Planck model for the probability of collapse with values c
1
= 0.01 and
c
2
= 2.64
still provide a useful functional dependence of the P
C
with the external and system parameters.
Clearly, a more realistic model for an earthquake motion should be nonstationary.
Further research is demanded in order to nd closed form solutions of the constants "c
1
" and
"c
2
" in terms of the system parameters.
8.5. Summary 215
8.5 Summary
Two expressions are proposed for the probability of collapse of rocking blocks under
earthquake loading. One is based on a modied version of the Poisson process. The
other has been developed based on the stationary solution of the associated Fokker-Plank
equation.
The Fokker-Plank based expression, although has been found by assuming restrictive
hypotheses, may still provide a useful functional dependence of the P
C
with the external
and system parameters.
Further research is needed to establish analytical expressions of the constants as functions
of the system parameters in the derived probability of collapse.
Part V
Conclusions and Future Research
216
9.1. Conclusions 217
9.1 Conclusions
Scope and Limitations
It has been shown both experimentally ([80]) and numerically ([58]) how masonry structures
can sustain acceleration well in excess of the limit implied by their static capabilities. This
fact has been the main motivation of the present contribution. In this regard, it has been found
through dierent perspectives that collapse is not equal to initiation of motion. Furthermore,
conditions for stability have been addressed for dierent types of loads and parameters.
On the other hand, the present study has not considered important eects which may take place
in the block dynamics:
1. No sliding has been considered.
2. Bouncing and other states implying deformation have been neglected.
3. 3D motion is only addressed in the theoretical formulation of multiblock assemblies, but
it has not been considered neither experimentally nor in the stability analysis.
With these limitations, the present work can only address an approximation to the motion of
simple planar congurations of block assemblies when no sliding is present.
Theoretical Development
RM Single block
1. The complex formulation presented in this work unies both piecewise equations of mo-
tion and impact mechanisms into a single model, in which the dynamics are governed by
a unique dierential equation.
2. Damping eects are understood by means of a generalized force constructed in a heuristic
manner. This force behaves as a Dirac-delta force, which is in agreement with the results
reported in previous works [50],[51].
3. The model proposed can be set in direct analogy with either a two-body central problem
in the complex plane or an inverted pendulum through simple variable transformations.
9.1. Conclusions 218
Multiblock analysis
The main results for multi-block dynamics are:
1. By geometrical methods, the formulation of single block dynamics has been extended to
an arbitrary number of blocks.
2. Results from rocking block dynamics can be used to assess simple masonry arch collapse
mechanisms and other closed kinematic chain structures.
3. The algorithms presented in this work used in combination with limit analysis routines
and DEM software can form a powerful procedure to tackle masonry structures dynamics.
Stability analysis
The stability analysis has been tackled through the following approaches:
Amplitude-frequency stability maps for harmonic forcing.
Bifurcation analysis.
Poincare surface of section methods.
Recurrence plots and recurrence quantication analysis.
Direct integration. Numerical curves for the probability of collapse.
The most important results are:
1. Dierent tools from nonlinear dynamics and chaos theory have been used to analyze
the stability of the RM dynamics. The underlying structure of the phase-space has been
studied intensively and stability boundaries for both harmonic and random loading have
been derived.
2. Several quantication procedures have been discussed and applied to the problem of rock-
ing blocks under random loading. Finally, an estimator for the quantication of chaos has
been proposed.
9.2. Future Research 219
Applications
The theoretical model developed in this work has been applied to a practical subject; the proba-
bility of collapse of single block structures under earthquake loading. For this task, the follow-
ing approach has been carried out:
1. Both experimental and numerical investigation has been carried out to determine the
functional relationships between the probability of collapse and the system parameters
involved.
2. Two possible models for the probability of collapse have been proposed. One based in
a modied version of a Poisson process and other based on the Fokker-Planck equation.
Although restrictive assumptions have been made on the solution of the FPE, the model
found can still provide useful criteria for the estimation of the probability of collapse
under general random loading.
3. Once the model of P
c
is ne-tuned, it can used as a reasonable estimator for the problem
of collapse of more complex structures with one degree of freedom, such as the four-
hinged masonry arch or other closed kinematic chain structures.
It was not the authors intention to substitute the existing numerical techniques for seismic anal-
ysis but to provide a suitable complement, in order to improve the knowledge on the dynamics
of masonry structures.
9.2 Future Research
The present contribution does not provide a solution for the probability of overturn of multiple-
degree-of freedomstructures, but it may still serve as an adequate basis for further developments
in this eld.
The analytical approach presented in this work is also unable to determine the collapse mech-
anisms of a given structure under earthquake loading. In this respect, the complex formulation
is a further-step to study the dynamics of a mechanism already materialized in the structure.
9.2. Future Research 220
Therefore, the present formulation depends on other existing techniques such as Limit analysis,
FEM, DEM, etc.
The main developments required in order to improve and to strength the present study are:
1. Investigation of a more precise closed-form solution for the probability of collapse.
2. Further development of instability criteria for multi-2D block congurations.
3. Use of Stochastic Melnikov analysis in order to nd analytical stability thresholds for the
probability of collapse model.
4. Consideration of sliding and bouncing.
5. Extension of the stability study to 3D models with the assistance of additional experimen-
tal research.
Part VI
Annexes
221
Appendix A
Periods of a Curved-Base Block
Equations of Motion
Consider a rectangular block with a curved base as the one shown in Fig. 1.4. This system is
referenced as CB. The center of rotation coincides with the center of curvature of the circle
segment of the base.
The main geometrical parameter of the CB system is given by:
y
cc
y
cm
(A.1)
as the distance between the initial heights of center of mass and curvature. The coordinates
of the center of mass (X
cm
, Y
cm
) can be expressed in terms of the coordinates of the center of
rotation (X
cc
, Y
cc
):
X
cm
= X
cc
+ sin() (A.2)
Y
cm
= Y
cc
cos() (A.3)
It can be easily shown that the dynamic functions are:
T =
1
2
M

R
c

2
+
1
2
(M
2
+ I
c
)

2
+ M(

X
cc
cos() +

Y
cc
sin())

,
U = Mg[1 cos()],
222
223
where I
c
is the moment of inertia about the center of mass. The rst term in the kinetic energy
is purely translational while the second is only rotational. The third term couples rotation and
translation.
To investigate rotations only the second term is kept.
T
ROT
=
1
2
(M
2
+ I
c
)

2
(A.4)
The moment of inertia of the system is simply I = M
2
+ I
c
.
Now U
CB
does not present a discontinuity at the origin and can be expanded into series about
= 0. Keeping only the terms of second order in :
U
1
2
(Mg)
2
(A.5)
It can be realized that this form of potential corresponds to an harmonic oscillator.
Periods
The general expression for the periods is:
T = 4
_

0
0
d
1
_
2
I
(E U)
(A.6)
Being E the mechanical energy and
0
the amplitude of the oscillation. At the returning points
of the motion =
0
the kinetic energy equals zero. Therefore, since E is a constant of motion:
E = U
0
=
1
2
Mg
2
0
(A.7)
By substituting this expression into Eq. A.6 it is found that:
T = 2
_
I
Mg
(A.8)
If the pendulum-equivalent period is labeled by T
p
= 2
_

g
, the preceding expression can be
recast in the form of a series expansion:
T = T
p
_
1 + T
p
(1 +
1
2
+ O()
2
) (A.9)
224
where is dened as the non-dimensional parameter =
I
c
M
. This expression does not depend
on the amplitude of the oscillation.
Appendix B
Mathematical Results and Formulae
This Annex summarizes the set of mathematical expressions and results which have been used
along this work.
EULER IDENTITY
Intensive use of:
e
i
= cos() + isin() (B.1)
is made along the whole work.
HERMITIAN INNER PRODUCT
The Hermitian product in the complex vector space is dened through:
< ds, ds >=

i, j
g
i j
dx
i
dx
j
(B.2)
where the metric tensor is a symmetric bilinear form such that the following properties hold:
< u, v >= < u, v >
< u, v >=

< u, v >
< u, v >=< v, u >

< u
1
+u
2
, v >=< u
1
, v > + < u
2
, v >
< u, u >> 0 u 0
(B.3)
225
226
EXPONENTIATION OF AN OPERATOR
The exponentiation of an operator is another operator dened through the series expansion:
e
A
=

n=0
A
n
n!
(B.4)
ROTATIONS
An innitesimal rotation along the axis n by a quantity (see Fig. B.1) is represented as:
R

= 1 + n

S (B.5)
where

S is the so-called generator of the rotation. Finite rotations by an angle are then built
x+x
x

u
3
u
2
u
1
e
1
e
2
e
3
R
O
n
Figure B.1: Denition of rotations about an axis
as:
R

= e
n

S
(B.6)
If the axis n is aligned with the n-frame basis vector u
3
as shown in Fig. B.1, the preceding
expression renders:
R

= e

(B.7)
227
where the matrix , in 3D renders:
=

0 1 0
1 0 0
0 0 0

(B.8)
whereas in 2D = i
2
and
2
is the Pauli matrix:

2
=

0 i
i 0

(B.9)
For further clarication, see Arfken [53].
DIRECT SUMS OF MATRICES
A B =

A O
pxp
O
pxp
B

(B.10)
SCALARPRODUCTINTERMS OF MATRICES FORTHEEUCLIDEAN
SPACE
< u, v >= UV
T
= V
T
U (B.11)
ACTION OF AN OPERATOR ON A VECTOR IN TERMS OF MATRI-
CES
If
y =

A(x) (B.12)
then
Y = XA (B.13)
MATRIX OF OPERATOR
< u,

A(v) >= UA
T
V
T
= VAU
T
(B.14)
228
PROPERTIES OF THE UNITARY ROTATION OPERATOR USED IN
THE TEXT
RR
T
= I
R

i
R
T

j
= R

i

j
(B.15)
EXTENDED VECTORS IN THE PRODUCT SPACE
A vector "v" belonging to the tensor product space E
1
E
2
is dened through:
v = (v
1
, v
2
) (B.16)
where v
1
E
1
and v
2
E
2
UNDERLINED VECTORS
Vectors labeled with an underline and with a superscript n represent a vector belonging to a
tensor product space which all its components except those of the n-subspace are zero.
x
n
= (

0, ...,

0, x
n
,

0, ...,

0) (B.17)
PROPERTIES OF THE "s" SYMBOLS
s
n
1
+ s
n
2
= 1
s
n
1
s
n
2
= sign(
n
)
s
n
1
sign(
n
) = s
n
1
s
n
2
sign(
n
) = s
n
2
s
n
1
s
n
1
= s
n
1
s
n
2
s
n
2
= s
n
2
s
n
1
s
n
2
= 0
(B.18)
Investigation about what type of algebra these quantities could form is left for future research.
Appendix C
Dirac-Delta Forces and Phase Dynamics
The aim of this Appendix is to show that, during an impact, the quantity
f
l
(r)
1

d
dr
(C.1)
behaves as a Dirac-delta function.
The Dirac-delta function is dened in this work by its assigned properties:
_
b
a
g(x)(x x
0
) =

g(x
0
) ; x
0
(a, b)
1
2
g(x
0
) ; x
0
= a, b
0 ; elsewhere
(C.2)
As commonly assumed, (x) will be approximated as the limit sequence of continuous functions

l
(x). Strictly, it must be recalled that the limit
lm
l0

l
(x) (C.3)
does not exist. Nevertheless, (x) may be treated consistently in the form (see, for instance,
Arfken, [53])
_
b
a
(x)g(x)dx = lm
l0
_
b
a

l
(x)g(x)dx (C.4)
Selecting a = r
min
, one obtains
lm
l0
_
b
r
min

l
(r r
min
)g(r)dr =
1
2
g(r
min
) (C.5)
229
230
From Eqs. 5.19- 5.21 and the denition of Eq. C.1, the following expression is obtained:
f
l
(r) =
l
r
_
2(E U)Ir
2
l
2
(C.6)
The key quantity to be computed is:
lm
l0
_
b
r
min
f
l
(r)g(r)dr (C.7)
being g(r) an arbitrary function and r
min
given by Eq. 5.22. Through a change of variable
y = r/r
min
, and, after some manipulation, the following expression is obtained
lm
l0
_ b
r
min
1
g(yr
min
)dy
y
_
(
EU(yr
min
)
EU
0
)y
2
1
(C.8)
Calculating the limit (note that b is large, but nite), the result is
g(0)

_

1
dy
y
_
y
2
1
(C.9)
This integral can be calculated in the complex plane taking into account that the integrand has
a simple pole at y = 0 and two branch points at y = 1 (see Fig. C), giving /2.
1 0 1
R
x
y

Figure C.1: Contour of integration for Eq. C.9


231
Therefore
lm
l0
_
b
r
min
f
l
(r)g(r)dr =
g(0)
2
(C.10)
Comparing the above expression with Eq. C.4, since r
min
and l go simultaneously to zero, it is
possible to obtain
lm
l0
_
b
0
f
l
(x)g(x)dx =
g(0)
2
(C.11)
lm
l0
_
b
0

l
(x)g(x)dx =
g(0)
2
(C.12)
From this behavior of f
l
(x) on the arbitrary function g(x), it results that f
l
(x) can be identied
with a -sequence;
f
l
(x)
l
(x) (C.13)
and it is justied to assign:
d
dr
=
l
(r) (C.14)
From this equation, the dispersion of the angle at every impact of the block with the founda-
tion can also be obtained, as
lm
l0
=
_

(x)dx = (C.15)
By using the property of the Delta function (where the sum is extended over all the zeros (t
j
) of
the function x(t), e.g. the times of impact):
(x(t)) =
_
j
(t t
j
)

x(t
j
)

(C.16)
and by manipulation of Eq. C.14, one obtains
lm
l0
(t) =

j
(t t
j
) (C.17)
where d/dt = (t) and (t) is the Heaviside step function.
This last equation is tested numerically by integration of system of Eqs. 5.50 for two values of
l, being the results given in Fig. C.2. Upon decreasing value of l, excellent results are found for
the phase .
232
Figure C.2: Numerical test for the phase in Eq. C.17 for = 10
5
and = 10
2
.
Appendix D
Derivation of the associated
Fokker-Planck Equation
Let the normalized energy e(t) = E/I p
2
be a randomvariable. The stochastic process: {e(t), t T}
generates the Markovian chain, e(t
0
), e(t
1
), . . . , e(t
n
), which is formed by the random amplitudes
of the energy taken at discrete time instances: t
0
, t
1
, . . . , t
n
. The cumulative distribution function
has the Markov property:
P[e
n+1

n+1
|e
n
=
n
, e
n1
=
n1
, . . . , e
0
=
0
] = P[e
n+1

n+1
|e
n
=
n
] (D.1)
Notice that the random variable is termed as "e", whereas its value is labeled as "".
The joint probability distribution function is dened as usual:
(|
0
; t, t
0
) =
P[e(t) |e(t
0

0
)]

(D.2)
The Reliability function can now be dened as:
R(t) =
_
e
c
0
d(|
0
; t, t
0
) (D.3)
where e
c
represents the threshold energy value for collapse. This is the probability that the block
remains with an energy lower than e
c
. Therefore, the conditional probability of collapse is:
P
C|RM
(t) = 1 R(t) (D.4)
233
234
The associated Fokker-Planck equation (FPE) for (|
0
; t, t
0
) is given by [74]:

t
=
1
2

2
(a)

2

(b)

(D.5)
where b(, t) and a(, t) are the so-called drift and diusion coecients dened by:
b(, t) = lm
t0
< [e(t + t) e(t)|e(t) = ] >
t
, (D.6)
and
a(, t) = lm
t0
< [(e(t + t) e(t))
2
|e(t) = ] >
t
, (D.7)
where < > is the expected value operator. No analytical solution has been found for Eq. D.5.
Several approaches have tackled the problem by using sophisticated techniques to numeri-
cal solve the FPE. Since through an appropriate scaling, the FPE can be transformed into
a Schroedinger equation, techniques derived from quantum mechanics, such as path-integral,
have been used [71],[77].
The present contribution is mainly focused in the possible practical advantage of the probabilis-
tic approach from the earthquake engineering point of view. Therefore, a closed-form solution,
though not entirely correct, is more desirable than the solution under particular conditions of
force and parameters.
Instead of trying to solve numerically the FPE, a much simpler, although as it will be shown,
powerful approach, is followed in this work. The eort focus on to the stationary solution of
the FPE, easily found as:

s
=
c
a
exp
_
2
_
e
0
d
b
a
_
(D.8)
where "c" is an arbitrary constant.
Using Eqs. D.3, D.4 and D.8, the conditional probability of collapse reads:
P
C|RM
= 1
_
e
c
0
de
c
a
exp
_
2
_
e
0
d
b
a
_
(D.9)
From Eq. 7.14, the energetic relation can be recast in the form of an stochastic It equation [74]:
e = t + W (D.10)
235
where,
= ln()
_
j
( r
j
)
2
(t t
j
)
W = a
g
cos() sin( r)
(D.11)
Drift coecient
It must be noticed that the term W is a coupling between the external driving acceleration and
the angular signum, expressed by cos(). The sign function is regarded as a random variable
itself with zero mean: < cos() >= 0.
On the other hand, the external acceleration is a zero mean process (see section 6.2). Therefore,
it is reasonable to assume: < W >= 0 and hence, < e >= t < >. The drift coecient is
then given by: b(, t) =< |e(t) = >.
The quantity
_
j
(t t
j
)( r
j
)
2
is only non-zero at impacts. Since, at impact times, r
2
= 2e the
following expression holds:
b(, t) = 2ln() < F(t) > . (D.12)
where the quantity, F(t) =
_
j
(t t
j
), has been introduced.
Diusion coecient
According to Eq. D.6 the diusion coecient is:
a(, t) = lm
0
_
(t) <
2
> +
1
t
< (W)
2
> +2 < W >
_
(D.13)
By assuming < W >= 0, the only surviving term is < (W)
2
>. By introducing the relative
angle: z = r, the diusion coecient can be expressed as:
a(, t) = lm
t0
_
1
t
< (a
g
)
2
z
2
>
_
(D.14)
It is now assumed that < (a
g
)
2
z
2
>=< (a
g
)
2
>< z
2
>. That is, the increment of the process
a
g
is assumed as independent of z. With this hypothesis:
a(, t) = lm
t0
_
1
t
< (a
g
)
2
>
_
< z
2
> (D.15)
236
On the other hand, by using that a
g
= a
g
t, the expression:
< (a
g
)
2
>= t
_

S
a
g
()d (D.16)
is obtained, where, S
a
g
() is the power spectral density of the process a
g
. Since it also holds
that, S
a
g
() =
2
S
a
g
() (see [74]), the diusion coecient renders:
a(, t) =< z
2
>
_

d
2
S
a
g
() (D.17)
The power spectral density can be integrated from Eq. 6.6.
Appendix E
Experimental Details
Data Acquisition systems
The uniaxial seismic table of LNEC has a service loading, only limited by the bearings inner
strength, of 6 tons. It holds a maximum displacement of 10 mm and it is able to develop
an acceleration of 4.5 g for an input frequency of 20 Hz. These dynamical features resulted
adequate for the type of experiments performed.
The measuring system was selected in order to enable accurate position measurement of the
specimens, at each instant of the tests, and at the same time to avoid the possibility that the
system inuences the response of the specimens. In this context, the image processing and
measuring system is based upon the monitoring of Light Emission Diode Systems (LEDS) by
means of high resolution cameras. The Position Sensor is a light spot position detector which
is able to measure two-dimensional positions of light spots. A semiconductor position-sensitive
detector (PSD) is used as the light receiving element. The PSD is a sensor using a photodiode.
A LED target is connected to the object being measured, and the position is measured optically.
This eliminates error caused by noise conductance which is a problem with other objects such
as ordinary vibration meters and accelerometers, and enables accurate position measurement.
The position resolution obtained in this way is 1/5000. The measurement values are obtained as
serial output to the data monitoring system accomplished through National Instruments devices
and associated software.
237
238
The main measures obtained with this system were: rotations Y and Z, and linear displacements
X and Y. The system coordinate is depicted in Figure E.1. Rotations were measured by means
Figure E.1: Reference system of the measuring system.
of a mirror linked to the blocks surface. The LEDS were located at the same position of the
measuring camera and the light ray emitted is reected at the mirror. Details of the mirror
system data acquaintance are shown in the left of Figure E.2 and in the right shows a general
view of a typical RM test. The linear displacements were obtained in two points of the North
Figure E.2: System data acquaintance. LED-Camera-Mirror system (left) and general view
(right)
face of each specimen, through the LEDS coordinates directly (Fig. E.3). After every test, the
239
specimens were set back to their initial positions, since small changes in the initial position
were observed in each test.
Additionally to the LED-Camera system, two biaxial accelerometers were placed at the top of
each specimen, one in the North face and the other one in the South face. The displacements
Figure E.3: Details of LED-Camera system. Specimen 2 (left) and 4 (right)
and accelerations of the seismic table were measured too.
Two security frames were placed at each side (East and West) of the specimen in order to protect
the table and the stones in case of overturning. Since in pure Rocking Motion regime the system
has only one degree of freedom (DOF), monitoring of extra DOF permits additional control of
possible out-of-plane and sliding motions.
Parameter tting
The method used in determining the dynamical parameters (p, , ) was based on the criteria
that the values adopted in the theoretical model would yield best matching to the corresponding
response observed in the experiment. The approach is the same than that used by Wong & Tso
[40].
240
The empirical values of and were determined by minimizing the error surface (, ),
dened by Eq. E.1 and by solving the corresponding system of equations given in Eq. E.2.
(, ) =
N

n=1
(T
n
T
exp
n
)
2
(E.1)

= 0 (E.2)

p
= 0
Denition of T
n
is given in Eq. E.3 and Fig. E.4 while T
exp
n
are the corresponding times obtained
from experiment. For slender blocks and small angles, Housner [7] derived an expression for
0
1
2
3

1
max

2
max
|

|
(
o
)
(s)
Definition of times and periods
r
0
r
1
r
2
T
1
T
2
T
3
Experimental amplitude
Figure E.4: Denition of T
n
is given in Eq. E.3
T
n
given by Eq. E.3
T
n
=
1
p
cosh
1

1
1
r
n1

(E.3)
where r
n1
represent the amplitudes before impact n takes place.
Experimental amplitudes (and their corresponding times) are found by setting the condition
241
|| = 0 and then nding r
n
by the bisection method.
The resulting system of Eqs. E.2 is numerically solved with the help of the command fsolve
available in Maple giving the experimental values of and p for each impact.
Regarding the coecient of restitution , from Eq. 2.16, the following equation is obtained:

n
= (
cos( r
n
) cos()
cos( r
0
) cos()
)
1
2n
(E.4)
In this expression r
n
represent the amplitudes of the rocking angle after the n
th
impact while r
0
is the starting amplitude.
For a given test, the values of
n
for each impact are computed and then, the mean value is
obtained. Those mean values are then averaged again through the whole set of tests so that a
nal value for each specimen is found. The result from this process is given in Table E.1.
Table E.1: Theoretical and experimental classic parameters.
Specimen (rad) p (1/s)
T E % T E % T E %
1 0.242 0.235 -2.9 0.914 0.936 2.4 3.78 3.84 1.6
2 0.168 0.163 -3.0 0.958 0.973 1.6 3.81 4.05 6.3
3 0.119 0.154 29.4 0.978 0.978 0.0 3.82 3.61 -5.5
4 0.310 0.268 -13.5 0.860 0.927 7.8 5.16 5.02 -2.7
T = Theoretical, E = Experimental, % = Error (in percentage)
Appendix F
Sources of Error and Their Eects
Inuence of the Shaking Table Noise
In order to study the inuence of the shaking of the seismic table response in the behavior of
the rigid blocks, the harmonic and random movements are analyzed. Figure F.1 a) shows the
input and output records of the shaking table for a typical constant sine (5.0 Hz of frequency
with amplitude of 5 mm).
From the transfer function shown in Fig. F.1 b) it is clear that the output signal is equal to the
input signal, except at the end of the loading. This is due to the fact that the input signal ends
abruptly and the table cannot reproduce this phenomenon, the impulsive nal deceleration of
the table introduces a mechanical noise. The presence of this mechanical noise is evident in
the transfer function between the output and input acceleration records. The transfer function
exhibits three peaks at 15, 20 and 25 Hz. The rst two peaks are associated to the initial
(acceleration) and nal (deceleration) table motion; while the 25 Hz peak is associated to the
natural frequency of the table. The presence of this natural frequency is solely due to the fact
of the seismic table being switched on, with independence of any input signal.
In the case of harmonic motion, the possible divergences in the response induced by the noise
are not evidenced through numerical simulation. The only part in which the mechanical noise
aects in an appreciable amount the response of the blocks is at the end of loading. In that
242
243
a) b)
Figure F.1: Typical response of the shaking table under constant sine (5 Hz with 5 mm of am-
plitude); a): displacements, b): Fourier spectra and transfer function.
part, the impulsive deceleration of the base modies the free rocking behavior of the block (see
Fig. F.2). As a consequence, the extra noise induced by the table can be neglected along the
Figure F.2: Typical response under harmonic motion (Specimen 2, constant sine with frequency
of 5 Hz and 5 mm of amplitude).
calculations for the case of harmonic loading.
Figure F.3 a) shows the input and output displacement records of the shaking table for a typical
random motion (earthquake 4). For this case, the shaking table accurately reproduces the input
244
a) b)
Figure F.3: Typical response of the shaking table under random motion (Earthquake record 3);
a) displacements, b) accelerations.
signal. On the other hand, Fig. F.3 b) shows the input and output accelerations. It can be noticed
that the response of the table displays a high frequency content due to the impulsive forces of
the table. The transfer function between the output and input accelerations (Fig. F.4 a)) shows
that both signals are equal in a range between 0.5 and 10 Hz.
a) b)
Figure F.4: Typical response of the shaking table under random motion (Earthquake record 3);
a) Fourier spectra and transfer function, b) response spectra of acceleration.
245
However, there is an amplication of the frequencies up to 10 Hz, corresponding to the frequen-
cies ascribed to the shaking table noise. Figure F.4 b) illustrates a typical response spectrum of
acceleration for input and output accelerations. The response of the table has a very high peak
at 0.5s (20 Hz) corresponding to the induced noise. It must be highlighted that this noise cannot
be ltered since, as stressed above, it is an inherent element of the process of acceleration itself.
Through numerical simulation it has been observed that this mechanical noise does not neces-
sarily induce rocking motion by itself, but it adds certain amount of energy to the system. This
extra energy turns the block more vulnerable to the base motion.
Eect of Three-Dimensional Behavior
Along the experimental research, out-of-plane mechanism was evidenced. This fact modied
the response of the blocks.
In order to minimize out-of-plane motion, the specimens were manufactured with a thick-
ness/width (d/b) ratio equal to 3. However, as it will be shown in this section, this value for d/b
is not sucient to guarantee perfect in-plane motion.
As it is shown in Fig. F.5, no sliding or rotation around the vertical axis was noticed for speci-
men 1. On the other hand, specimen 2 experienced a rotation around the vertical axis in every
test. For this specimen, only for tests without rocking, torsion was not evidenced. This fact
suggests that both behaviors may become linked under certain conditions. Fig. F.6 shows the
response of specimen 2 for a constant-amplitude sinusoidal load of 3 Hz of frequency and 8
mm of amplitude. The maximum torsion angle observed at the end of the testing is close to one
degree.
246
a) b)
Figure F.5: Specimen 1, hanning sine 3.3Hz and 10mm. a) horizontal displacements, b) rocking
and torsion angle
a) b)
Figure F.6: Specimen 2, constant sine 3Hz and 8mm. a) horizontal displacements, b) rocking
and torsion angle
It is noticed that, for harmonic load, the magnitude of the torsion angle is related to the ampli-
tude of the load and not to the frequency (Table F.1). However, for random motion, no clear
relationship between the torsion angle and the loads (Fig. F.7) was found. On the other hand,
specimen 4 presents a small rotation around vertical axis, as shown in Fig. F.8. This displace-
ment can only be ascribed to torsion eects and not to sliding. As it was the case for specimen
247
Table F.1: Specimen 2, maximum torsion angle (
o
) for harmonic motion.
Constant Sine Amplitude (mm)
(Hz) 03 04 05 06 08 10 12
1.5 0.88
2.0 NT NT NT 0.88 0.88
2.5 NT NT 0.52 0.33 0.88
3.0 0.14 0.30 0.55 0.64 0.95
3.3 0.14 0.15 0.55 0.74 0.70
5.0 0.74 0.10 0.64
Hanning Sine (Hz)
2.5 NT 0.10
3.0 0.10 0.15 0.30
3.3 0.12 0.25
NT = No Torsion
Figure F.7: Specimen 2, maximum torsion angle for random motion
248
a) b)
Figure F.8: Specimen 4, run-down sine 7.0 0.7 Hz and 2 mm. a) horizontal displacements,
b) rocking and torsion angle
2, the maximum torsion angle depends on the amplitude of the load and not on its frequency
(Table F.2). Specimen 3 displays a very dierent behavior than the other specimens. In addition
Table F.2: Specimen 4, maximum torsion angle (
o
) for harmonic motion.
Constant Sine Amplitude (mm)
(Hz) 02 03 04 05
4.0 0.12 0.18 0.12
6.5 0.13 0.22 0.38
Runup 0.77.0 0.12 0.23
Rundown 7.00.7 0.36
to torsional eects, this specimen exhibits a vibration around the vertical axis. Figure F.9 shows
the angles and horizontal displacements corresponding to a constant sine of 5 Hz and 2 mm.
That gure evidences the vibration around vertical axis.
It is also noticed that the horizontal displacements are related to the vibration and not to the slid-
ing motion. This vibration around the Y axis has the eect of modifying the aspect ratio of the
block, turning the specimen into a block with a "larger" basis. This is illustrated in Fig. F.10.
249
a) b)
Figure F.9: Specimen 3, constant sine 5Hz and 2mm. a) horizontal displacements, b) rocking
and torsion angle
Since the vibration frequency can be considered as signicant when compared with the typical
RM "frequency", the base dimension of the vibrating block results in an averaged width which
is over its theoretical value. The subsequent eect is that the parameter values are heavily af-
fected, as inspection to table E.1 states. Specimen 1 is the only one which presents a perfect
Figure F.10: Specimen 3. Vibration around Y axis.
in-plane behavior whereas stones 2 and 4 exhibit a little rotation around Y axis. However, this
eect is particularly evident in specimen 3.
250
It is suggested that the inertia moments ratio (I
0
/J
0
) could be a good parameter for such analy-
sis. The inertia moments of the specimens are summarized in Table F.3. It is clear that specimen
3 has the lowest value of torsional inertia, whereas specimen 1 has the greatest one.
However, the study of these rotational eects is out of the scope of the present work. The fo-
cus on reporting such eects is to evidence those possible causes that could lead to a loss of
agreement between observed and theoretical parameters.
Table F.3: Geometrical characteristics of the specimens.
Specimen M 2b 2h 2d I
0
J
0
I
0
/J
0
(Kg) (m) (m) (m) (Kg-m
2
) (Kg-m
2
)
1 506 0.25 1.00 0.754 179 26.6 6.7
2 228 0.17 1.00 0.502 78 5.3 14.7
3 122 0.12 1.00 0.375 42.8 1.6 26.8
4 245 0.16 0.457 0.750 26.2 12.0 2.2
Appendix G
Computer Programs
General Methodology
In this section, the computer programs used along this work are summarized. The procedures
can be ordered according to their use:
Numerical Integration.
Symbolic Manipulation.
Data Analysis.
Numerical integration algorithms
Integration routines are c-code based in the GSL (Gnu Scientic Library) libraries under the
GPL license. The GSL is available on the World Wide Web at [81].
The GSL algorithms allow great exibility in the choice of integration schemes. Several al-
gorithms were tested and the most adequate one turned out to be the type "rkf", based on the
fourth-fth order Runge-Kutta-Fehlberg method. The GSL library "gsl-odeiv-control" allowed
accurate control on the convergence of the solution at each time step. Then "gsl-odeiv-evolve"
was used to integrate the dierential equations forward in time.
251
252
Symbolic Manipulation Software
The formulation developed in Part II deals with analytical expressions. Symbolic manipulation
of mathematical formulae is an essential task in the context of multiblock dynamics. This task
is aided by c Maple 9 based algorithms. Here, the main procedures are briey described.
The Maple-based codes MBLOCK2D.mws and MBLOCK3D.mws were designed to interface
with Unix Bash scripts in order to integrate both the numerical analysis and the symbolic
manipulation. MBLOCK2D.mws and MBLOCK3D.mws use the intermediate routine MRE-
DUCE.mws, which implements the constraints from Eqs. 3.79 in order to reduce the initial
degrees of freedom. The global MREDUCE.mws procedure is as follows:
1. Input geometry. Dene, node positions and weights.
2. Input fracture lines and hinge positions.
3. Derivation of d and x
c
vectors for each mechanism.
4. Computation of operators (matrices) [C, P, ].
5. Calculus of rotations R.
6. Derivation of generalized tensors: I, A, B, Z
c
, Q, J, F for each mechanism.
7. Computation of the mechanism tensors S

.
8. Coupling between rotations and mechanisms. Derivation of the intermediate tensors K
and Y.
9. Build-up of the potential and kinetic energy expressions.
10. Decomplexication.
Data Analysis Procedures
PSS
The Poincar surface of section is computed with the program poincare.c based on the GSL
routines. On the other hand, visual inspection of the phase space structure is aided with the
program c Phaser.
253
Recurrence Analysis
Recurrence plots are performed via the TISEAN package "recurr" [66]. Quantication is ac-
complished through the Commandline recurrence plots, version 1.13z (2006/03/09 12:00:00)
2005-2006 from the "Nonlinear Dynamics Group", available at the World Wide Web [73].
Bibliography
[1] Solares J.M. Los efectos en Espaa del terremoto de Lisboa (1 de Noviembre de 1755).
Instituto Geogrco Nacional, 2001.
[2] Pelaez J.A. and Lopez Casado C. Seismic hazard estimate at the iberian peninsula. Pure
and Applied Geophysics, 159:26992713, 2002.
[3] Ministerio de Fomento. Norma de construccin sismorresistente (NCSE-02). Parte general
y de edicacin, 2002.
[4] Eurocode 8: Design of structures for earthquake resistance. Part 1; general rules, seismic
actions and rules for buildings, 2005. EN 1998-1.
[5] FEMA-HAZUS. http://www.fema.gov/hazus.html.
[6] Venice Charter. http://www.icomos.org/venicecharter.html.
[7] Housner W. G. The behavior of inverted pendulum structures during earthquakes. Bulletin
of the Seismological Society of America, 53:403417, 1963.
[8] Makris N. and Konstantinidis D. The rocking spectrum and the limitations of practical
design methods. Earthquake Engineering and Structural Dynamics, 32:265289, 2003.
[9] Lipscombe P. R. Dynamics of rigid block structures. PhD thesis, University of Cambridge,
Cambridge, England, 1990.
254
BIBLIOGRAPHY 255
[10] Loureno P. B., Rots J. G., and Blaauwendraad J. Continuum model for masonry: Param-
eter estimation and validation. Journal of Structural Engineering, ASCE, 124(6):642652,
1998.
[11] Heyman J. The Stone Skeleton. Cambridge University Press, 1995.
[12] Hogan S. J. On the dynamics of rigid block motion under harmonic forcing. Proceedings
of the Royal Society of London A, 425:441476, 1989.
[13] Hogan S. J. Heteroclinic bifurcations in damped rigid block motion. Proceedings of the
Royal Society of London A, 439:155162, 1992.
[14] Spanos P. D. and Koh A. M. Rocking motion of rigid blocks due to harmonic shaking.
Journal of Engineering Mechanics, 110, 1984.
[15] Yim C. S. and Lin. H. Nonlinear impact and chaotic response of slender rocking objects.
Journal of Engineering Mechanics, 117(9):20792099, 1991.
[16] Zhang J. and Makris N. Rocking response of free-standing blocks under cycloidal pulses.
Journal of Engineering Mechanics, ASCE, 127:473483, 2001.
[17] Augusti G. and Sinopoli A. Modelling the dynamics of large block structures. Meccanica,
27:195211, 1992.
[18] Eurocode 8: Design of structures for earthquake resistance. Part 3: Assessment and
retrotting of buildings, 2005. EN 1998-3.
[19] Meli R. Structural engineering of historical buildings (in Spanish). ICA, 1998.
[20] Livesley R. K. Limit analysis of structures formed fromrigid blocks. International Journal
of Numerical Methods in Engineering, 12:18531871, 1978.
[21] Gilbert M. and Melbourne C. Rigid-block analysis of masonry structures. The Structural
Engineer, 72(21):356361, 1994.
BIBLIOGRAPHY 256
[22] Begg D. W. and Fishwick R. J. Numerical analysis of rigid block structures including
sliding. Proceedings of the 3rd International Symposium of Computational Methods in
Structural Mechanics, pages 177183, 1995.
[23] Baggio C. and Trovalusci P. Limit analysis for no-tension and frictional three-dimensional
discrete systems. Mechanics of Structures and Machines, 26(3):287304, 1998.
[24] Ordua A. and Loureno P. B. Three-dimensional limit analysis of rigid block assem-
blages. Part I: Torsion failure on frictional interfaces and limit analysis formulation. In-
ternational Journal of Solids and Structures, 42:51405160, 2005.
[25] Ordua A. and Loureno P. B. Three-dimensional limit analysis of rigid block assem-
blages. Part II: Load-path following solution procedure and validation. International Jour-
nal of Solids and Structures, 42:51615180, 2005.
[26] Ferris M. C. and Tin-Loi F. Limit analysis of frictional block assemblies as a mathematical
program with complementarity constraints. International Journal of Mechanical Sciences,
43:209224, 2001.
[27] Loureno P. B. and Rots J. G. A multi-surface interface model for the analysis of masonry
structures. Journal of Engineering Mechanics, ASCE, 123(7):660668, 1997.
[28] Psycharis N., Lemos J. V., Papastamatiou D. Y., Zambas C., and Papantonopolous C.
Numerical study of the seismic behaviour of a part of the parthenon pronaos. Earthquake
Engineering and Structural Dynamics, 32:20632084, 2003.
[29] Winkler T., Meguro K., and Yamazaki F. Response of rigid body assemblies to dynamics
excitation. Earthquake Engineering and Structural Dynamics, 24:13891408, 1995.
[30] Pea F., Prieto F., Loureno P. B., Costa A. C., and Lemos J. V. On the dynamics of
rocking motion of single rigid-block structures. Sent for possible publication, 2007.
[31] Milne J. Experiments in observational seismology. Transactions of the Seismological
Society of Japan, 3, 1881.
BIBLIOGRAPHY 257
[32] Winkler T., Meguro K., and Yamazaki F. Response of rigid body assemblies to dynamic
excitation. Earthquake Engineering and Structural Dynamics, 24:13891408, 1995.
[33] Psycharis I. N. Dynamic behavior of rocking two-block assemblies. Earthquake Engi-
neering and Structural Dynamics, 19:555575, 1990.
[34] A. Sinopoli. Nonlinear dynamic analysis of multiblock structures. Structural Dynamics
(eds. W. B. Kratzig et al.), 1:244259, 1991.
[35] Doherty K., Grith M., Lam N., and Wilson J. Displacement-based seismic analysis
for out-of-plane bending of unreinforced masonry walls. Earthquake Engineering and
Structural Dynamics, 31:833850, 2002.
[36] De Felice G. Out-of-plane fragility of historic masonry walls. In Taylor and London
Francis Group, editors, Structural Analysis of Historical Constructions, pages 11431148.
Modena, Loureno and Roca, 2005.
[37] Koon W. S. and Marsden J. E. The hamiltonian and lagrangian approaches to the dynamics
of nonholonomic systems. Reports on Mathematical Physics, 40:2162, 1997.
[38] Leon M. and Martin de Diego D. On the geometry of non-holonomic lagrangian systems.
Journal of Physics, 37(7):33843414, 1996.
[39] Aslam M., Godden W. G., and Scalise D. T. Earthquake rocking response of rigid bodies.
Journal of the Structural Division. Proceedings of the American Society of Civil Engineer-
ing, 106(2):378392, 1980.
[40] Wong C. M. and Tso W. K. Steady state rocking response of rigid blocks Part 2: Experi-
ment. Earthquake Engineering and Structural Dynamics, 18:107120, 1989.
[41] Yim C. S., Chopra A. K., and Penzien J. Rocking response of rigid blocks to earthquakes.
Earthquake Engineering and Structural Dynamics, 8:565587, 1980.
[42] N. Makris and Y. S. Roussos. Rocking response of rigid blocks under near-source ground
motions. Geotchnique, 50(3):243262, 1999.
BIBLIOGRAPHY 258
[43] Liberatore D. Rocking of slender blocks subjected to seismic motion of the base. In Pro-
ceedings of the 12th European Conference on Earthquake Engineering, 2002. Ref(760).
[44] Ishiyama Y. Motions of rigid bodies and criteria for overturning by earthquake excitations.
Earthquake Engineering and Structural Dynamics, 10:635690, 1982.
[45] Liberman M. A. and Lichtenberg A. J. Regular and Chaotic Dynamics. Springer-Verlag,
1992.
[46] Hogan S. J. Slender rigid block motion. Journal of Engineering Mechanics, 120(1):1124,
1992.
[47] Yim C. S. and Lin. H. Chaotic behavior and stability of free-standing oshore equipment.
Ocean Engineering, 18(3):225250, 1991.
[48] Shenton H. Criteria for initiation of slide, rock, and slide-rock rigid-body modes. Journal
of Engineering Mechanics, ASCE, 122(7):690693, 1996.
[49] Jeong M. Y., Suzuki K., and Yim S. Chaotic rocking behavior of free-standing objects
with sliding motion. Journal of Sound and Vibration, 262:10911112, 2003.
[50] Prieto F., Loureno P. B., and Oliveira C. S. Impulsive dirac-delta forces in the rocking
motion. Earthquake Engineering and Structural Dynamics, 33:839857, 2004.
[51] Prieto F. and Loureno P. B. On the rocking behavior of rigid objects. Meccanica,
40(2):121133, 2005.
[52] Goldstein H. Classical Mechanics. .Addison-Wesley, Reading, Massachusetts, 1950.
[53] Arfken G. B. and Weber H. J. Mathematical Methods for Physicists. Academic Press,
fourth edition, 1995.
[54] Spence R. and Coburn A. Strengthening buildings of stone masonry to resist earthquakes.
Meccanica, 27:213221, 1992.
BIBLIOGRAPHY 259
[55] Lipscombe P. R. and Pellegrino S. Free rocking of prismatic blocks. Journal of Engineer-
ing Mechanics, 119(7):13871409, 1993.
[56] Spanos P. D., Roussis C. P., and Politis P. A. Dynamic analysis of stacked rigid blocks.
Soil Dynamics and Earthquake Engineering, 21:559578, 2001.
[57] Kovacs-Bende M. Response of rigid block assemblies to earthquake excitation. Technical
Report 1, Servian Polytechnical University, 2000.
[58] Grith M. C., Magenes G., Melis G., and Picchi L. Evaluation of out-of-plane stability
of unreinforced masonry walls subjected to seismic excitation. Journal of Earthquake
Engineering, 7(1):141169, 2003.
[59] Psycharis I. N., Papastamatiou D. Y., and Alexandris A. P. Parametric investigation of the
stability of classical columns under harmonic forcing and earthquake excitations. Earth-
quake Engineering and Structural Dynamics, 29:10931109, 2000.
[60] Transgranitos, web page. http://www.transgranitos.pt/pt.html.
[61] Makris N. and Roussos Y. Rocking response and overturning of equipment under horizon-
tal pulse-type motions. Technical report, University of California, Berkeley, 1998. PEER
Report 1998/05.
[62] Pompei A., Scalia A., and Sumbatyan M. A. Dynamics of rigid block due to horizontal
ground motion. Journal of Engineering Mechanics, 124(7):713717, 1998.
[63] Azevedo J, Sincrain G., and Lemos J. V. Seismic behavior of blocky masonry structures.
Earthquake Spectra, 16(2):337365, 2000.
[64] Tso W. K. and Wong C. M. Steady state rocking response of rigid blocks. Part 1; Analysis.
Part 2: Experiment. Earthquake Engineering and Structural Dynamics, 18:89120, 1989.
[65] Marwan N. Encounters with neighbors. Current developments of concepts based on re-
currence plots and their applications. PhD thesis, University of Postdam, 2003.
BIBLIOGRAPHY 260
[66] Hegger R., Kantz H., and Schreiber T. Practical implementation of nonlinear time series
methods: the TISEAN package. Chaos, 9:413435, 1999.
[67] Lenci S. and Rega G. A dynamical systems approach to the overturning of rocking blocks.
Chaos, Solitons and Fractals, 28:527542, 2006.
[68] Ott E. Chaos in Dynamical Systems. Cambridge University Press, 1993.
[69] Hilborn R. C. Chaos and Nonlinear Dynamics. An Introduction for Scientists and Engi-
neers. 2nd Edition. Oxford University Press, 2000.
[70] Eckmann J. and Ruelle D. Recurrence plots of dynamical systems. Europhysics Letters,
5:973977, 1987.
[71] Lin H. and Yim C. S. Nonlinear rocking motions. II: Overturning under random excita-
tions. Journal of Engineering Mechanics, 122(8):728735, 1996.
[72] Zbilut J. P. and Webber Jr. C. L. Embeddings and delays as derived from quantication of
recurrence plots. Physics Letters A, 171:199203, 1992.
[73] Commandline recurrence plots, version 1.13z (2006/03/09 12:00:00) c 2005-2006.
[74] Solnes J. Stochastic Processes and Random Vibrations. Theory and Practice. Wiley &
Sons, 1997.
[75] Giannini R. and Masiani R. Non-gaussian solution for random rocking of slender rigid
block. Probabilistic Engineering Mechanics, 11:8796, 1996.
[76] Facchini L., Gusella V., and Spinelli P. Block random rocking and seismic vulnerability
estimation. Engineering Structures, 16(6):412424, 1994.
[77] Cai G. Q., Yu J. S., and Lin Y. K. Toppling of rigid block under evolutionary random base
excitations. Journal of Engineering Mechanics, 121(8):924928, 1995.
BIBLIOGRAPHY 261
[78] Iyengar R.N. and Manohar C.S. Rocking response of rectangular rigid blocks under ran-
dom noise base excitation. International Journal of Non-linear Mechanics, 26:885892,
1991.
[79] Augusti G., Baratta A., and Casciatti F. Probabilistic Methods in Structural Engineering.
1st ed. Chapman and Hall, London, 1984.
[80] NTUA. Monuments under seismic action. a numerical and experimental approach. Techni-
cal report, Laboratory for Earthquake Engineering. Faculty of Civil Engineering. National
Technical University of Athens, 1997. Report No. NTUA/LEE-97/01.
[81] Gnu Scientic Library. http://www.gnu.org/software/gsl/.

You might also like