You are on page 1of 110

I.U.S.S.

Istituto Universitario Universit degli


di Studi Superiori Studi di Pavia

EUROPEAN SCHOOL OF ADVANCED STUDIES IN


REDUCTION OF SEISMIC RISK
ROSE SCHOOL

SEISMIC VULNERABILILTY OF MASONRY


ARCH BRIDGE WALLS

A Dissertation Submitted in Partial


Fulfilment of the Requirements for the Master Degree in

EARTHQUAKE ENGINEERING

By

MARIA ROTA

Supervisors: Prof. ALAIN PECKER


Dr. RUI PINHO

May, 2004
The dissertation entitled Seismic vulnerability of masonry arch bridge walls, by Maria Rota,
has been approved in partial fulfilment of the requirements for the Master Degree in Earthquake
Engineering.

Rui Pinho

Alain Pecker
ACKNOWLEDGEMENTS

The author would like to thank Dr. Rui Pinho for his guidance and support during the work
and for his useful critique during the final stages of the dissertation.

The author is also grateful to Prof. Alain Pecker for his valuable help and encouragement
during the period spent in Paris, and for his careful review of the manuscript.

Last but not least, the important contribution of Mr. Davide Bolognini is acknowledged, for his
insightful comments and suggestions.

3
Table of contents

TABLE OF CONTENTS

1 INTRODUCTION.......................................................................................................................... 10
2 ARCH MASONRY BRIDGES ..................................................................................................... 13
2.1 General characteristics of masonry bridges ......................................................................... 13
2.2 Characteristics of the filling material..................................................................................... 14
2.3 Seismic damage to masonry bridges...................................................................................... 15
3 SEISMIC OUT-OF-PLANE CAPACITY OF WALLS............................................................ 19
3.1 Introduction.............................................................................................................................. 19
3.2 Out-of-plane capacity of isolated walls................................................................................. 19
3.3 Out-of-plane capacity of typical masonry arch bridge walls ............................................. 26
4 STATIC SOIL PRESSURES ON A RETAINING WALL ..................................................... 30
4.1 Introduction.............................................................................................................................. 30
4.2 Static soil pressure ................................................................................................................... 30
4.3 Check for sliding failure of the wall ...................................................................................... 31
4.4 Check for overturning failure of the wall............................................................................. 32
4.5 Check for shear failure of the wall ........................................................................................ 33
5 DYNAMIC SOIL PRESSURES ON A RETAINING WALL ............................................... 36
5.1 State of the art .......................................................................................................................... 36
5.2 Method by Mononobe and Okabe........................................................................................ 37
5.2.1 Description of the model .............................................................................................. 37
5.2.2 Shortcomings of the model........................................................................................... 41
5.3 Method by Scott (1973) .......................................................................................................... 41
5.3.1 General derivation of the model .................................................................................. 41
5.3.2 Discussion of some details of the model .................................................................... 44
a) Boundary effects.................................................................................................................. 44
b) Effect of the wall flexibility................................................................................................ 45
5.3.3 Shortcomings of the model........................................................................................... 45
5.4 Method by Veletsos and Younan [1994b]............................................................................ 46
5.4.1 Description of the model .............................................................................................. 46
5.4.2 Shortcomings of the model........................................................................................... 52
5.4.3 Further developments of the method [Veletsos and Younan, 1997 and 2000]..... 53

4
Table of contents

5.5 Comparison of the results obtained with the three methods implemented.................... 53
6 COMBINED MODEL COMPUTATION OF CAPACITY............................................... 58
6.1 Introduction.............................................................................................................................. 58
6.2 Parametric study on the reciprocal influence of the two walls ......................................... 59
6.3 Overview of capacity model................................................................................................... 62
6.3.1 Introduction..................................................................................................................... 62
6.3.2 Determination of the acceleration capacity of the wall............................................. 62
6.4 Calibration of the parameters of Veletsos and Younan [1994b] method........................ 65
6.4.1 Stiffness of the rotational spring .................................................................................. 65
6.4.2 Reduction of shear modulus ......................................................................................... 69
6.4.3 Frequency of excitation ................................................................................................. 71
6.4.4 Discussion on the influence of acceleration on the results ...................................... 73
6.5 Results obtained ....................................................................................................................... 77
7 COMPUTATION OF DEMAND ............................................................................................... 83
7.1 Introduction.............................................................................................................................. 83
7.2 Amplification of the seismic input to the base of the wall ................................................ 83
7.3 Amplification of the seismic input through the wall .......................................................... 85
7.4 Results obtained ....................................................................................................................... 86
8 COMPARISON OF CAPACITY AND DEMAND................................................................. 94
9 CONCLUSIONS AND FUTURE DEVELOPMENTS........................................................ 105
10 REFERENCES .............................................................................................................................. 107

5
List of figures

LIST OF FIGURES

Fig. 2.1: identification of the different parts constituting a masonry arch bridge [Galasco et al.,
2004] ................................................................................................................................................... 13
Fig. 2.2: typical thickness of the filling material..................................................................................... 15
Fig. 2.3: overturning of the walls for a railway masonry bridge, after the ......................................... 16
Fig. 2.4: clear evidence of overturning of a bridge wall, after Koyna earthquake, India, 1967 [ASC]
............................................................................................................................................................. 17
Fig. 2.5: overturning of the wall of a masonry bridge, after the Umbria-Marche earthquake, 1997
............................................................................................................................................................. 18
Fig. 3.1: constitutive law used for masonry............................................................................................ 20
Fig. 3.2: forces acting on the wall ............................................................................................................ 21
Fig. 3.3: stress diagrams on the masonry section at the different significant points of the
curve. The tensile stresses are not represented, since the masonry does not resist in tension.
Note that this is the same procedure used for a reinforced concrete section.......................... 22
Fig. 3.4: example of - curve ................................................................................................................ 25
Fig. 3.5: maximum acceleration capacity obtained for the different walls considered .................... 28
Fig. 3.6: values of maximum acceleration obtained for walls of different slenderness h/t ............ 28
Fig. 3.7: values of maximum acceleration obtained for walls with different axial load P ............... 29
Fig. 4.1: factors of safety obtained against sliding (CS), for the different wall cases considered.... 32
Fig. 4.2: factors of safety obtained against overturning (CO), for the different wall cases
considered .......................................................................................................................................... 33
Fig. 4.3: factors of safety obtained against shear (CV), for the different wall cases considered ..... 35
Fig. 5.1: forces acting on an active wedge in M-O analysis, for a dry cohesionless backfill (from Wood
[1973]) .................................................................................................................................................. 39
Fig. 5.2: system considered by Scott [1973] ........................................................................................... 42
Fig. 5.3: System considered by Veletsos and Younan [1994b]............................................................ 47
Fig. 5.4: Model of the soil stratum: elastically constrained bar ........................................................... 49
Fig. 5.5: comparison of base shear, obtained with the two methods, for different values of R ... 55
Fig. 5.6: comparison of the points of application of the thrust resultant, obtained with the two
methods, for different values of R ................................................................................................ 56
Fig. 5.7: comparison of base moment, obtained with the two methods, for different values of R
............................................................................................................................................................. 56

6
List of figures

Fig. 6.1: graphical representation of a complex quantity, c .................................................................. 59


Fig. 6.2: system considered by Wood [1973] ......................................................................................... 60
Fig. 6.3: dimensionless thrust (left) and moment (right) factors......................................................... 60
Fig. 6.4: base shears and base moments obtained for different values of L/H ............................... 61
Fig. 6.5: forces acting on the wall in the combined model .................................................................. 63
Fig. 6.6: pressure distribution on the wall for different values of rotational stiffness R................ 66
Fig. 6.7: effect of wall flexibility on the base shear, for a statically excited system ...................... 67
Fig. 6.8: effect of wall flexibility on the base moment, for a statically excited system ................ 67
Fig. 6.9: stiffness of the rotational spring for the different cases considered ................................... 68
Fig. 6.10: shear modulus reduction curve [Seed et al., 1986]............................................................... 70
Fig. 6.11: reduction of the shear modulus for the different cases considered, for = 300........... 71
Fig. 6.12: variation of the soil pressure distribution for different values of , for a fixed-base wall
(left) and for a rotationally constraint wall, with R = 1.00 E7 (right)...................................... 72
Fig. 6.13: variation of G/Gmax, hVel and max, with the level of acceleration A, for case 7 and =
400....................................................................................................................................................... 75
Fig. 6.14: variation of G/Gmax, hVel and max, with the level of acceleration A, for case 4 and =
100....................................................................................................................................................... 76
Fig. 6.15: capacity of the walls considered, for 5 different bridge typologies ................................... 78
Fig. 6.16: variation of the base shear with the ratio frequency of excitation - first frequency of the
wall, for different values of wall flexibility d [Veletsos and Younan, 1994b] ......................... 79
Fig. 6.17: variation of the first frequency of the soil layer, with the frequency of excitation ......... 80
Fig. 6.18: variation of the ratio /1 with the height of the soil layer .............................................. 80
Fig. 6.19: capacity of the first 7 cases of wall considered..................................................................... 81
Fig. 6.20: variation of the acceleration capacity with the height of the wall, h.................................. 82
Fig. 6.21: acceleration capacity of the different wall cases obtained with and without considering
the effect of the infill material......................................................................................................... 82
Fig. 7.1: example of elastic horizontal response spectra, for seismic zone 4 (PGA = 0.05 g) ....... 84
Fig. 7.2: acceleration response factor for a damping ratio = 0.05................................................... 85
Fig. 7.3: acceleration response factors for the case of = 300 .......................................................... 86
Fig. 7.4: variation of the acceleration demand with the soil category, for the different bridge
typologies and the different walls considered............................................................................... 89

7
List of figures

Fig. 7.5: variation of the acceleration demand with the seismic zone, for the different bridge
typologies and the different walls considered............................................................................... 91
Fig. 7.6: variation of the acceleration demand with the seismic zone and the different soil
categories, for the wall of case 12................................................................................................... 92
Fig. 7.7: variation of the acceleration demand with the bridge typology, for the wall of case 17 .. 93
Fig. 8.1: comparison of acceleration capacity and demand, for all the bridges and walls cases
considered, assumed to be on soil of type B, C, and E and in seismic zone 2........................ 96
Fig. 8.2: ratio between demand and capacity obtained for the different bridge typologies ............ 97
Fig. 8.3: ratio between demand and capacity obtained for the different bridge typologies and for
the different soil types...................................................................................................................... 99
Fig. 8.4: ratio between demand and capacity, for the different bridge typologies and seismic zones
...........................................................................................................................................................101
Fig. 8.5: variation of acceleration demand and capacity with bridge typology, for the wall of case 9
and for each seismic zone..............................................................................................................103
Fig. 8.6: variation of acceleration demand and capacity with bridge typology, for the wall of case 9
and for each soil type .....................................................................................................................104

8
List of tables

LIST OF TABLES

Table 2.1: values of specific weight for some types.............................................................................. 15


Table 3.1: characteristics of the walls considered in the parametric study. ....................................... 26
Table 5.1: characteristics of the system considered .............................................................................. 54
Table 5.2: results obtained from the comparison of Veletsos and Younan and M-O methods, for
the case of fixed massless wall and static excitation ................................................................ 54
Table 5.3: results obtained with the three method for a fixed-base wall, excited with a frequency
of 400 rad/s ....................................................................................................................................... 57
Table 6.1: percentage variation of base shear and base moment........................................................ 62
Table 6.2: characteristics of the system considered .............................................................................. 72

9
Chapter 1 Introduction

1 INTRODUCTION

Many railway and road bridges in Italy (and, more generally, in Europe) are arch masonry
bridges. It is a very old typology of bridges, which employment can be traced back to the ancient
Romans period, or even earlier. The durability of this kind of bridges is remarkable, as
demonstrated by the many roman masonry arch bridges, as well as some aqueducts, which are
still perfectly working. This construction technique has been transmitted from one generation to
the other, with only minor variations. Recently, new bridges are more often steel or concrete
structures, but many masonry bridges still exist and their seismic performance is a major concern.
The potentially high seismic vulnerability of masonry railway bridges has not yet been fully
perceived, maybe because there have not been many significant damage evidences after recent
earthquakes. While this can be simply due to chance, it is a matter of fact that it is hard to find
any specific research study on this subject in the literature.
In particular, the problem of the interaction between the infill material and the side walls of the
bridges under seismic excitation has not been studied yet, but appears to be a potentially relevant
damage mode. Above all, a problem to be considered is the interaction between infill material
and parapet walls, in relation with a possible out-of-plane collapse. This failure mode is expected
to occur at the very beginning of the seismic excitation, before other types of mechanisms may
have damaged the structure. Also, it is believed that this kind of out-of-plane collapse may occur
even for low levels of acceleration. This mechanism of failure has been studied in the current
work, through a parametric study on a set of bridges and walls typologies. For each case,
acceleration demand and capacity are compared and some conclusions are drawn.

The present work is organised in 9 Chapters, which will be briefly described here. Following the
introduction, Chapter 2 is an overview of the main characteristics of arch masonry bridges,
particularly railway bridges, that will be studied in this work. The different parts composing such
a structure are identified and special attention is given to the characteristics of the filling material,
introduced between the two walls of the bridge, in order to create a horizontal plane for the rails.
Also, an overview of typical seismic damages, occurred in the past to this typology of bridges, is
presented. It is shown that local failures, such as overturning of the bridge walls, probably due to
the interaction with the infill material, are very common for masonry arch bridges.

10
Chapter 1 Introduction

In order to study whether this failure mechanism occurs or not for a given bridge typology, a
comparison between capacity and demand is carried on, for a wide range of cases. Attention is
given particularly to the determination of the capacity of the walls, including the effect of the
infill material. For this scope, a capacity model is presented, obtained from the combination of
different sub-models, each one describing a single aspect of the problem. In a first moment, the
different sub-models are introduced separately and, only in a second time, they are assembled, to
obtain the combined model, as explained more in detail in what follows.
In Chapter 3, a method developed by Priestley [1985] and Paulay and Priestley [1992] for the
out-of-plane seismic capacity of an isolated wall is introduced. Using this force-based model, it is
possible to determine the acceleration-displacement curve of an unreinforced masonry wall and
therefore to evaluate the maximum acceleration that the wall can resist before collapse. The
described method is applied to a set of typical arch masonry walls and the results obtained are
presented.
After the maximum acceleration that a bridge wall can resist is determined, the thrust of the
infill material on the walls needs to be taken into account, since it influences the capacity of the
wall. This thrust can be subdivided into a static and a dynamic part. The static thrust, which has
been calculated using Coulombs theory, is described in Chapter 4. Then, the typical bridge walls
previously introduced are checked with respect to three static mechanisms of failure, i.e. sliding,
overturning and shear.
The calculation of the dynamic component of the infill thrust is discussed in detail in Chapter 5.
Three different methods for the evaluation of this dynamic thrust are implemented in the current
work, respectively proposed by Mononobe-Okabe, Scott and Veletsos and Younan. The results
obtained with these three methods are reported and compared. At the end, it is shown that the
method of Veletsos and Younan is preferable in the current framework.
The model for the isolated wall (described in Chapter 3) and the two models for the static and
dynamic components of the thrust (described in Chapters 4 and 5), are combined, in order to
form a complete capacity model for the system consisting in the walls and the infill material. This
combined model is described in detail in Chapter 6. In order to apply this model, it is necessary
to calibrate some parameters used in Veletsos and Younan method, to determine the dynamic
thrust. These are essentially the stiffness of the rotational spring at the base of the wall, the
frequency of excitation and the reduction of the shear modulus of the soil, depending on the
level of deformation. Moreover, some other details of Veletsos method, concerning the
maximum acceleration at the base of the wall, need to be analysed. Each one of these aspects is

11
Chapter 1 Introduction

described in detail in Chapter 6. The acceleration capacity, obtained for each one of the wall cases
considered, after application of this combined model, is also reported and discussed.
Chapter 7 is devoted to the computation of the level of demand on the walls considered. This
demand is calculated in two different steps: in a first moment, the acceleration demand at the
base of the bridge wall is determined, using an elastic response spectrum. This spectrum is
constructed according to the new Italian Seismic Code [O.P.C.M. n. 3274, 2003] and the
acceleration corresponding to the dominant frequency of oscillation of the bridge in the
transverse direction is read from the spectrum. In particular, the acceleration at the base of the
wall for all the wall cases considered is determined, assuming that the bridge belongs to each one
of the four seismic zones described in the Code (each one corresponding to a different level of
peak ground acceleration on rock) and assuming it is located on each one of the three soil
categories. Then, the acceleration demand at the barycentre of the wall is determined, after
application of an acceleration response factor. The results obtained are also reported and
discussed in the same Chapter.
Chapter 8 summarises the most relevant results obtained from the comparison of the
acceleration capacity and demand for the different cases. In particular, the ratio between demand
and capacity is analysed, since it is believed to be of some importance not only to check if a wall
fails or survives, but also to quantify how it is far from the limit condition, corresponding to
demand equal to capacity.
Finally, Chapter 9 summarises the most relevant conclusions derived from this work and
outlines suggestions for future research, with particular emphasis for those issues where further
investigation is required.

12
Chapter 2 Arch masonry bridges

2 ARCH MASONRY BRIDGES

2.1 General characteristics of masonry bridges


As stated by Resemini [2003], the modern Italian masonry bridges, especially railway bridges,
have been built in a time period of 100 years, approximately between 1830 and 1930. This very
short period determines the homogeneity of some of the construction techniques and of the
geometry of such structures. Nevertheless, they show some different details, depending on the
year of construction, on the geographical area and, most likely, also on the designer. For a more
detailed overview of the history of masonry arch bridges construction, the reader is referred to
the work of Resemini [2003].
Generally, the parts constituting a masonry bridge (shown in Fig. 2.1) are: the arch, which is the
structural part of the bridge; the elements supporting the arch, i.e. abutments and piers; the
foundations and the non-structural parts, located above the arch, to create a horizontal plane,
such as the filling material. The filling material is laterally constrained by two walls, which are
located above the exterior part of the arch. Some more detail on the characteristics of this infill
material will be discussed in the next section.

Fig. 2.1: identification of the different parts constituting a masonry arch bridge [Galasco et al., 2004]

13
Chapter 2 Arch masonry bridges

For existing masonry bridges, one of the major difficulties consists in the determination of the
characteristics of the materials used in the construction. This is mainly due to the intrinsic
characteristics of masonry, which is a very anisotropic material, whose mechanical properties are
strongly dependent on the properties of its constituents. In particular, for existing structures it is
far from trivial to determine the consistency of the grout, when present, or the quality of the
structural elements (bricks or stones) used. Often, different types of materials may be used for
different parts of the bridge, due to structural reasons (need for higher resistance of the more
heavily stressed parts and for lower weight of the non-structural parts), as well as to economical
reasons. [Resemini, 2003]
The geometry of the bridge is strongly influenced by the orography of the valley to be crossed.
Wide and deep valleys are often crossed by bridges called viaducts, with more spans on high
piers, whilst wide but shallow valleys are crossed by bridges with more spans on short piers.
Narrow valleys and minor streams are usually crossed by single-span bridges. [Resemini, 2003]

2.2 Characteristics of the filling material


The space between the two walls of a masonry arch bridge is filled up with some material, in
order to create a horizontal plane. This infill material must be light and able to drain the water;
moreover it should contribute to the repartition to the arch of the concentrated loads applied on
the horizontal plane on top of it.
The filling material is usually incoherent material (such as soil or mucking resulting from mines
excavation) or, in order to reduce the thrust on the walls, it may be constituted by dry stones,
coarse aggregate, gravel, ballast or, more recently, low resistance concrete. In case of viaducts,
especially when they have piers of significant height, it is not uncommon to find some brick
vaults instead of the filling material. The reasoning behind this solution is not clear, but it may
have been adopted to reduce the weight acting on the arch. For the sake of simplicity, this infill
material will be always referred to as soil, in what follows.
Some typical values of the specific weight of the infill material, suggested by Gambarotta et al.
[2001b], are summarised in Table 2.1. In that work, it is also stated that, when precise
information on the type of material used is lacking, a value between 17 and 19 KN/m3 may be
reasonably assumed for the specific weight. In the present work a value of 17 KN/m3 has been
selected.

14
Chapter 2 Arch masonry bridges

Table 2.1: values of specific weight for some types


of filling material
Material Specific weight [kN/m3]
Incoherent material 16 18
Dry stones 18 21
Aggregate or gravel 14.5 18
Low resistance concrete 21

According to Albenga [1953], for railway bridges, the height of the soil between the horizontal
plane and the top of the arch must not be less than 40 cm. However, for lower-height bridges,
this limit may be brought down to 30 cm, but never less than 15 cm. Generally, the thickness of
the stratum is equal to the arch thickness at the apex stone, as illustrated in Fig. 2.2.

Fig. 2.2: typical thickness of the filling material

The filling material produces, as obvious, some pressures on the side walls of the bridges. This
pressure can be subdivided into static (always present) and dynamic (developed only when the
structure is subjected to dynamic loads) parts. The calculation of the static part is described in
detail in the chapter 4, whilst the dynamic part will be described in chapter 5.

2.3 Seismic damage to masonry bridges


As stated above, most of the existing Italian masonry bridges have been built between 1830 and
1930. Therefore their seismic history is quite short, with a consequent lack of information about
seismic damages to these structures. It seems likely that, for earthquakes of moderate intensity, a
masonry bridge, without particular structural deficiencies, will survive without being heavily
damaged; nevertheless, the response of these bridges to major earthquakes requires more
investigation [Resemini, 2003].

15
Chapter 2 Arch masonry bridges

Information about seismic damage to masonry bridges throughout the world is very limited. In
fact, in the less industrial areas of the world, masonry structures are either of limited extension, or
information is difficult to be obtained, whilst in the more industrial zones, modern infrastructures
are not realized in masonry, as anticipated in the introduction. Also, the seismic risk of this kind
of structures is not associated with human lives loss, but with the functionality of some parts of
the system of infrastructures. This is another reason of the scarce information available, especially
for those areas in which an earthquake causes many victims and therefore this kind of
infrastructural damage is not considered of significance. Therefore, the seismic vulnerability of
masonry bridges seems to be an interesting issue only in Europe and, in particular, considering
the seismicity of the area, in Italy [Resemini, 2003].
Even if the construction techniques may vary from area to area, due to local knowledge and in
situ availability of materials, a study of the evidences of seismic damage to masonry bridges
throughout the world may help in the knowledge of the damage mechanisms of these structures.
The Bhuj earthquake (magnitude 7.7), which occurred in January 2001 in the district of Gujarat,
India, caused damage to several masonry bridges. Most of them were short span bridges, with
short piers. The damages which have been observed consist in cracking and skews of the
elements of the arches, damages to the piers and overturning of the bridge walls. An example of a
railway masonry bridge, which has been damaged during the Bhuj earthquake, is shown in Fig.
2.3. This 88 years old bridge is already under repair, as indicated in the picture. It can be seen
that the earthquake caused the overturning of the walls and the falling-off of the infill material,
exposing the train rails.

Fig. 2.3: overturning of the walls for a railway masonry bridge, after the
Bhuj earthquake, 2001, India [Gisdevelopment, 2001]

16
Chapter 2 Arch masonry bridges

Another example of seismic damage to a masonry bridge is due to the earthquake of Koyna,
which occurred in India (1967) and is shown in Fig. 2.4. The damage clearly consists in the
overturning of one of the bridge walls. The infill material between the two walls, which is visible
in the picture, consists in some kind of soil.

Fig. 2.4: clear evidence of overturning of a bridge wall, after Koyna earthquake, India, 1967 [ASC]

Another example of damage due to overturning of the bridge walls is related to the earthquake
occurred in the Italian regions of Umbria and Marche, in 1997. The damaged arch masonry
bridge is shown in Fig. 2.5. Also in this case, the overturning of the wall exposes the filling
material, even if, from this picture, it is not easy to identify the type of material constituting the
filling for this particular bridge.

17
Chapter 2 Arch masonry bridges

Fig. 2.5: overturning of the wall of a masonry bridge, after the Umbria-Marche earthquake, 1997
[Resemini and Lagomarsino, 2004]

The above puts in evidence that one of the most frequent types of failure consists in the
overturning of the bridge walls. For these bridges, local failures, such as overturning of the bridge
walls, probably due to the interaction with the infill material, are very common. The present work
will focus on the study of this particular failure mechanism.

18
Chapter 3 Seismic out-of-plane capacity of walls

3 SEISMIC OUT-OF-PLANE CAPACITY OF WALLS

3.1 Introduction
The determination of the capacity of unreinforced masonry walls to out-of-plane seismic
excitation is one of the most complex and ill-understood areas of seismic analysis [Griffith et al,
2003]. One of the reasons for this is the fact that all modern codes for new design of masonry
structures give dimensioning and detailing rules that make out-of-plane collapse of walls
extremely unlikely. Nevertheless, out-of-plane collapse is a crucial issue for old existing masonry
structures [DAyala and Speranza, 2002].
There exist different types of out-of-plane collapse mechanisms that can be observed, being
related to the overall geometry of the wall, its boundary conditions and the quality of the
materials [DAyala and Speranza, 2002; Lagomarsino, 1999]. In this work, the attention is
restricted to the simplest mechanism, the one way bending condition, which occurs in cantilever
walls, since this is the typology of restraint typical of masonry arch bridge walls. In this case, the
formation of cracks does not constitute wall failure, even in unreinforced walls. Failure can occur
only when the resultant force in the compression zone of the central crack, R, is displaced
outside the line of action of the applied gravity loads, as will be discussed later in further details.
A simplified model is used to determine the acceleration-displacement curve and therefore to
obtain the maximum level of acceleration that a given wall can resist prior to collapse. This model
has been applied to a set of walls, which are considered typical for masonry arch bridges, and the
maximum values of acceleration that each wall can resist, without considering the effect of the
infill material, have been calculated. The results obtained are summarised in section 3.3.

3.2 Out-of-plane capacity of isolated walls


In this work, the approach developed by Priestley [1985] and Paulay and Priestley [1992] for a
masonry wall is followed, in order to analytically evaluate the capacity of a wall; in particular, the
acceleration-displacement curve is determined. The wall is modelled as a single-degree-of-
freedom structure and is subjected to a seismic action in the direction perpendicular to the wall.
The following assumptions are involved in the method:
The response acceleration is constant over the height of the wall. Therefore the lateral
inertia force per unit area will be w = ma, where m is the mass per unit area of wall

19
Chapter 3 Seismic out-of-plane capacity of walls

surface. This force can also be written as w = p, where p is the weight per unit area of
wall surface and is the seismic coefficient, given by = a/g.
The model considers a wall of constant, rectangular section and unitary width.
P - effects are included in the model.
The constitutive relationship for the material is nonlinear, for which reason the elasto-
perfectly-plastic behaviour shown in Fig. 3.1 has been assumed for the masonry.

f max

Fig. 3.1: constitutive law used for masonry

Failure of the wall will occur for out-of-plane deflection, after the formation of a plastic
hinge at the base. In particular, referring to Fig. 3.2, the wall will fail when the point of
application of the resultant of the vertical loads (P and W) is aligned with the point of
application of the reaction R.
The model can consider the presence of an axial load P, acting on top of the wall (in
case of arch masonry bridge walls there could be a parapet on top of each wall).
The method conservatively assumes that the displacement at the top of the wall
increases proportionally to the curvature. A more accurate estimate of the displacement
could be obtained by integrating the curvature distribution, but this is probably not
warranted, since the errors deriving from this approximation are typically not large,
until very large displacements are reached [Paulay and Priestley, 1992].
The model was originally developed by Paulay and Priestley for two different cases of
boundary conditions, i.e. for a wall fixed at the base and free at the top and for a wall
simply supported at both the ends. In this work, for obvious reasons, only the first
condition has been considered.

20
Chapter 3 Seismic out-of-plane capacity of walls

P
P

p /2
h
W

h/2

o
R
x

Fig. 3.2: forces acting on the wall

The seismic coefficient = a/g is considered in this model as a load factor and is used to
construct the - curve. This curve will be referred in what follows as the acceleration-
displacement curve, since is effectively a dimensionless acceleration, normalised by a constant
parameter (g).
The acceleration-displacement curve is obtained by sequentially evaluating its points, increasing
progressively the curvature at the top section of the wall and imposing equilibrium of external
and internal forces. The forces acting on the wall are represented in Fig. 3.2. The force indicated
as P is the inertia force produced by the axial load P. This force is proportional to P through the
seismic coefficient .
For a generic value of eccentricity x of the reaction R, the seismic coefficient can be
determined through the equation of equilibrium of moments around the point O in Fig. 3.2:

p h 2 (3.1)
Rx = + P h + P + W
2 2

from which:

W
Rx P +
= 2 (3.2)
ph
h + P
2

21
Chapter 3 Seismic out-of-plane capacity of walls

As stated above, the - curve needs to be defined by a series of discrete points. This is
because the relationship between acceleration and displacement is influenced by the nonlinear
constitutive law of the material and cracking effects.
The first point of the - curve to be determined is the point of first cracking. Up to this point,
the section is intact and therefore the - curve is linear elastic. At this point, the stress diagram
is triangular, as can be seen from Fig. 3.3, and therefore the maximum eccentricity x is given by
t/6, if t is the thickness of the wall. The cracking point is particularly important for this method,
since the determination of the displacements is based on the values of displacement and
curvature at cracking, as will be shown later in this section.

origin f=R/t

cracking f=2R/t

2R/t<f<f max
after cracking

f=f max
yielding
b1

f=f max
after yielding
b

f=f max
ultimate
t

Fig. 3.3: stress diagrams on the masonry section at the different significant points of the curve. The
tensile stresses are not represented, since the masonry does not resist in tension. Note that this is the same
procedure used for a reinforced concrete section

The seismic coefficient at cracking, indicated as *, can be obtained by calculating the


overturning moment from equilibrium considerations and by setting it equal to Rt/6 (the area of
the triangular stress distribution in Fig. 3.3). It should be pointed out that, at this stage, the

22
Chapter 3 Seismic out-of-plane capacity of walls

second order moment due to P and W is negligible and therefore it is not included in the
expression for *:

t
R
* = 6 (3.3)
ph
h + P
2

Before cracking, the wall behaviour is linear elastic and therefore the displacement can be
calculated applying the superposition of the effects due to the inertia force uniformly distributed
along the wall (p) and to the inertia force concentrated at the top (P). From classic structural
mechanics formulae, the displacement at cracking * is thus given by:

1 ph 4 1 * Ph 3 (3.4)
* = * +
8 EI 3 EI

where E is the elastic modulus of masonry and I is the moment of inertia of the intact section.
The corresponding curvature * can be obtained dividing the moment at cracking (Mcrack) by EI,
i.e.:

M crack R t (3.5)
* = =
EI 6 EI

After first cracking has occurred, the process becomes nonlinear. The crack propagates through
the section and the stress diagram remains triangular until the maximum resistance, fmax, of the
masonry is reached. This situation corresponds to the point of first yielding, which is the next
significant point of the - curve.
In this range of the curve, is calculated according to Eq. (3.2). The stress f in the masonry is
given by:

2R
f =
t (3.6)
3 x
2

23
Chapter 3 Seismic out-of-plane capacity of walls

The curvature can be obtained from:

f 2R
= =
t 2
(3.7)
3 E x 3 t x E
2 2

With the assumption of displacement at the top proportional to the curvature, can be
calculated as:

* (3.8)
=
*

After yielding, the stress diagram becomes trapezoidal, as shown in Fig. 3.3. For a given value
of b, b1 can be calculated as:

2R
b1 = b (3.9)
f max

and the eccentricity x can be obtained from:

1 2
(b b1 ) b + b1 + b12 (3.10)
t
x= 3 3
2 b + b1

The curvature is given by:

f max
= (3.11)
E (b b1 )

The seismic coefficient and the displacement are obtained respectively according to Eq. (3.2)
and Eq. (3.8).
Finally, the last point of the curve to be determined is the ultimate point, corresponding to the
failure of the wall. The ultimate displacement u occurs when the resisting moment becomes

24
Chapter 3 Seismic out-of-plane capacity of walls

zero, i.e. when the reaction R is aligned to the resultant of the vertical loads. Thus, from moment
equilibrium:

P +W (3.12)
u = xmax where:
P +W / 2

t a 1 R
xmax = = t (3.13)
2 2 2 f max

In the previous equation, a represents the width of the rectangular stress block at ultimate
conditions. As expected, to the ultimate displacement corresponds a seismic coefficient = 0.
An example of - curve, obtained with the method described above, is reported in Fig. 3.4,
just to give an idea of the shape of this kind of curves.

0.3

0.25 max

0.2
= a/g

0.15

0.1

0.05
u
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

Displacement [m]

Fig. 3.4: example of - curve

A couple of comments to the described model must be made:


The assumption that the displacement at the top of the wall is proportional to the
curvature is not really accurate. Hence, the values of displacement derived with this
method are not accurate. Nevertheless, in this work, attention is paid more to the value
of maximum acceleration that a wall can withstand than to the corresponding
displacement. Therefore, it is believed that the model is accurate enough for the
purposes of this study.

25
Chapter 3 Seismic out-of-plane capacity of walls

According to Paulay and Priestley [1992], in calculating the response using the
methodology outlined above, some account of vertical acceleration should be taken,
since this reduces the equivalent acceleration necessary to induce failure in the wall.
Conservatively, a value equal to two-thirds of the peak lateral ground acceleration is
suggested by Paulay and Priestley. However, the vertical acceleration contribution has
been neglected in this work.

3.3 Out-of-plane capacity of typical masonry arch bridge walls


A parametric study has been performed, considering a set of walls of different dimensions and
characteristics, which are summarised in Table 3.1. The values of the geometrical parameters
have been chosen based on realistic considerations on the dimensions of existing bridges and
based on limitations due to the selected method for the calculation of the dynamic thrust of the
filling material on the wall, which will be discussed later. These issues will be analysed in some
detail in the next chapters.

Table 3.1: characteristics of the walls considered in the parametric study.


Parameters common to all cases: fmax = 10MPa, E = 800 fmax
Case h [m] t [m] P/W Case h [m] t [m] P/W
1 1 0.4 0 15 1.7 0.8 0
2 1 0.4 0.5 16 1.7 0.8 0.5
3 1 0.5 0 17 2 0.8 0
4 1 0.5 0.5 18 2 0.8 0.5
5 1.2 0.5 0 19 2 1 0
6 1.2 0.5 0.5 20 2 1 0.5
7 1.2 0.6 0 21 2.2 0.9 0
8 1.2 0.6 0.5 22 2.2 0.9 0.5
9 1.5 0.6 0 23 2.2 1.1 0
10 1.5 0.6 0.5 24 2.2 1.1 0.5
11 1.5 0.7 0 25 2.5 1 0
12 1.5 0.7 0.5 26 2.5 1 0.5
13 1.7 0.7 0 27 2.5 1.2 0
14 1.7 0.7 0.5 28 2.5 1.2 0.5

Values of slenderness (h/t) of the walls approximately equal to 2 and 2.5, typical for this kind of
walls, have been used. In particular, the height is assumed to vary between 1 and 2.5 m, while the

26
Chapter 3 Seismic out-of-plane capacity of walls

thickness ranges from 0.4 to 1.2 m. These significant values of thickness are also confirmed by
the observations of Torre [2003], who states that a thickness value of 1 m is quite typical for the
walls of a masonry arch bridge and the range of variation of the thickness may be between 0.8
and 1.5 m, according to the type of infill material and therefore to the thrust acting on the walls.
It should be noted that the actual height of the walls of arch bridges varies along the arch. In
this study, a constant height has been assumed, ideally equal to the weighted mean height of the
wall.
The eventual presence of a parapet on top of the walls is considered in some cases, throughout
the application of an axial load P. This parapet would not in any case be very heavy and therefore
an axial load equal to half of the weight of the wall has been assumed.
Concerning the resistance of the masonry, a maximum value of 10 MPa has been considered.
This value comes from the considerations reported by Gambarotta et al. [2001a] about realistic
values of resistance of the different types of masonry used in the past for the construction of
arch bridges. In particular, Gambarotta and co-researchers suggest mean values of resistance
ranging from 10 to 30 MPa for brick masonry and from 5 to 30 MPa for stone masonry. From
these ranges, the assumption of a resistance of 10 MPa seems to be a conservative value and was
thus adopted in this work. The Young modulus E has been assumed to be 800 times the masonry
resistance, with 800 being a reasonable value based on previous studies on arch masonry bridges
[Picchi, 2001]. A masonry unit weight of 18 kN/m3 has been used to compute the wall weight.
The model described in the previous section for the determination of the acceleration-
displacement curve of an isolated wall has been applied to the selected cases (refer to Table 3.1).
For each case, the maximum acceleration max that the wall can withstand before failing has been
determined. The values obtained for the different cases are summarised in Fig. 3.5. It can be
observed that the maximum acceleration the walls can survive ranges from a minimum value of
0.285 g, to a maximum value of 0.488 g.

27
Chapter 3 Seismic out-of-plane capacity of walls

Maximum acceleration capacity w/o soil


0.6

0.5

0.4
max

0.3

0.2

0.1

0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

cases considered

Fig. 3.5: maximum acceleration capacity obtained for the different walls considered

As expected, the more slender the wall is (i.e. the higher is the ratio h/t), the lower the value of
max it can resist, as shown in Fig. 3.6. In this plot, walls of the same height and with the same
axial load P = 0, but with different slenderness values are compared. In particular, the maximum
accelerations obtained for walls with slenderness values approximately equal to 2 and 2.5 are
compared.

0.6
h/t = 2
0.5 h/t = 2.5

0.4
max

0.3

0.2

0.1

0
1 1.2 1.5 1.7 2 2.2 2.5
Height of the walls [m]

Fig. 3.6: values of maximum acceleration obtained for walls of different slenderness h/t

Also, the walls with an axial load on top can resist lower levels of acceleration than the walls
with the same geometry, but without axial load, as shown in Fig. 3.7. This effect is due to the

28
Chapter 3 Seismic out-of-plane capacity of walls

destabilizing effect of P and of the horizontal inertia force P induced by the axial load. In Fig.
3.7, only the walls with a slenderness value approximately equal to 2.5 are reported. If we look for
example at the first couple of cases in the plot, corresponding to walls of height h = 1 m , the
decrease of maximum acceleration due to the axial load is in the order of 26%.

0.5 P=0
P = 0.5
0.4

0.3
max

0.2

0.1

0
1 1.2 1.5 1.7 2 2.2 2.5
Height of the walls [m]

Fig. 3.7: values of maximum acceleration obtained for walls with different axial load P

29
Chapter 4 Static soil pressures on a retaining wall

4 STATIC SOIL PRESSURES ON A RETAINING WALL

4.1 Introduction
In the previous chapter, a simplified model for the behaviour of an isolated wall has been set
up. It is now necessary to consider the pressure that the infill material exerts on the wall. It has
already been anticipated that this effect can be subdivided into a static component and a dynamic
component. The static thrust of the soil on the walls, which is more straightforward, will be
discussed in this chapter, whilst the dynamic thrust will be analysed in greater detail in chapter 5.
Moreover, before calculating the effect of the dynamic pressure induced by the presence of the
soil, the out-of-plane stability of the walls under the static thrust has been checked, in order to
make sure that the selected parameters (geometry and characteristics of the walls) are realistic. In
particular, the static resistance of the walls has been checked with respect to three possible failure
mechanisms: sliding, overturning and shear. It should be underlined that the first failure
mechanism, sliding, concerns the overall stability of the wall, whilst the two others, overturning
and shear, are checked at the section level.

4.2 Static soil pressure


The static thrust exerted on a wall by the infill soil has been calculated using Coulombs theory
[e.g. Kramer, 1996]. According to this theory, the soil thrust is determined from force
equilibrium, for both active and passive conditions. The basic assumption is that the force acting
on the back of a retaining wall results from the weight of a wedge of soil, limited by a planar
failure surface. A significant number of failure surfaces should be analysed, in order to identify
the critical failure surface, i.e. the surface that produces the greatest active thrust or the smallest
passive thrust. It is evident that in the present work, the situation of interest is the active case,
corresponding to the soil exerting a force on the wall.
The active soil thrust on a wall retaining a cohesionless soil can be determined using the
following formulae:

1 (4.1)
Rst = K A h2
2

where:

30
Chapter 4 Static soil pressures on a retaining wall

cos 2 ( )
KA = 2
sin( + ) sin( i ) (4.2)
cos cos( + ) 1 +
2

cos( + ) cos(i )

In the above formulae:


is the unit weight of the soil
h is the height of the soil
i is the angle of inclination from horizontal of the backfill
is the angle of inclination from vertical of the wall
is the friction angle of the soil
is the angle of interface friction between the wall and the soil

Coulomb theory does not give explicitly the distribution of active pressure, but it can be
assumed that, for linear backfill surfaces without surface loads, this distribution is triangular. In
such cases, the static pressure Rst acts at a height of h/3, with h being the height of the wall and
this is what has been assumed in the current work.

4.3 Check for sliding failure of the wall


A sliding type of failure occurs when horizontal force equilibrium is not maintained, i.e. when
the lateral pressures on the back of the wall produce a thrust that exceeds the available sliding
resistance at the base of the wall. The sliding resistance at the base of the wall is given by the
product of the vertical forces times the tangent of the angle of interface friction. Therefore, the
sliding check is satisfied when CS > 1, with CS defined by Eq. (4.3):

CS =
(W + P + R )tan
st ,v
(4.3)
Rst ,h

where Rst,v and Rst,h are the vertical and horizontal components of the thrust, W is the weight of
the wall and P is the axial load applied at the top of the wall.
The results obtained for CS (see Table 3.1 for the identification of the cases considered), are
summarised in Fig. 4.1. It can be observed that, for all the cases considered, the factor of safety

31
Chapter 4 Static soil pressures on a retaining wall

against sliding is well above the limit condition, corresponding to CS = 1. For the soil density, a
value of = 1700 Kg/m3 has been used throughout the entire work.

6
5
factor of safety

4
3
2
1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Fig. 4.1: factors of safety obtained against sliding (CS), for the different wall cases considered

4.4 Check for overturning failure of the wall


Overturning moment failures occur when moment equilibrium is not satisfied, i.e. when the
overturning moment, caused by the thrust, exceeds the resisting moment, due to the vertical
forces acting on the wall. It is important to point out that all the calculations are made per linear
meter of wall. In particular, the resisting moment is given by:

t a (4.4)
M res = R
2 2

where t is the thickness of the wall and R is given by the sum of the vertical forces acting, i.e.:

R = W + P + Rst ,v (4.5)

with W being the weight of the wall and P being the eventual axial force acting on top of the
wall (both for an out-of-plane length of 1 m). Finally the term a in Eq. (4.4) is the width of the
compressed zone at the ultimate condition (when the maximum resistance of the wall is reached),
given by:

32
Chapter 4 Static soil pressures on a retaining wall

R
a= (4.6)
f max

where fmax is the maximum resistance of the masonry.


The overturning check is then satisfied when CO > 1, with CO defined by:

t a
R
CO = 2 2 (4.7)
h
Rst ,h
3

The results obtained for CO for the different cases considered (see Table 3.1), are summarised
in Fig. 4.2. It can be observed that, for all the cases considered, also the factor of safety against
overturning is well above the limit condition, corresponding to CO = 1.

8
7
factor of safety

6
5
4
3
2
1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Fig. 4.2: factors of safety obtained against overturning (CO), for the different wall cases considered

4.5 Check for shear failure of the wall


Shear failure occurs when the shear resistance Vt of the section is exceeded by the shear force
acting on the wall, which is the horizontal component of the static soil thrust, Rst,h. Again, all the
calculations are made per linear meter of wall. The shear check is satisfied when CV > 1, with CV
defined as:

Vt
CV = (4.8)
Rst ,h

33
Chapter 4 Static soil pressures on a retaining wall

In the equation above, Vt is calculated according to:

t ' f vk
Vt = (4.9)
M

where t is the thickness of the compressed zone of the wall, defined as :

t' = t with (4.10)

6M
1 if 1
tN
3 3 M 6M
(4.11)
= if > 1
2 t N tN
0 2 M
if 1
tN

where
- M is the bending moment around the axis perpendicular to the plane of the
figure, induced by the earth thrust: M = Rst,h*h/3 + Rst,v*t/2
- t is the thickness of the wall
- N is the axial force acting, given by the weight of the wall, W, plus the axial
force P, due to the eventual parapet on top of the wall
M is a safety factor, used only when designing new structures and therefore assumed
equal to one in this work, consisting in an assessment of the vulnerability of existing
bridges.
fvk is the shear resistance of masonry, defined as:

f vk = f vk 0 + 0.4 m 1.5MPa 1.4 f bk (4.12)

where
- fvk0 is the characteristic value of the shear resistance, when no compression is
applied to the section

34
Chapter 4 Static soil pressures on a retaining wall

- m is the average normal tension, calculated in this case as m = N/t


- fbk is the characteristic value of the vertical compression resistance of masonry
blocks

The results obtained for CV for the different cases considered (see Table 3.1), are summarised
in Fig. 4.3. It can be observed that, for all the cases considered, also the factor of safety against
shear is well above the limit condition, corresponding to CV = 1.

80
70
factor of safety

60
50
40
30
20
10
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Fig. 4.3: factors of safety obtained against shear (CV), for the different wall cases considered

35
Chapter 5 Dynamic soil pressures on a retaining wall

5 DYNAMIC SOIL PRESSURES ON A RETAINING WALL

5.1 State of the art


Despite the multitude of studies that have been carried out over the years, the dynamic
response of vertical retaining walls is still not well understood.
The methods that have been used until now can be classified into three categories:
(1) limit state analyses, in which the relative motions of the wall and backfill material are
sufficiently large to induce in the soil a limit or failure state;
(2) methods in which the wall and soil movements are limited, so that it can be assumed that
the backfill material deforms in its linear elastic range. This approach is often referred to
as the elastic approach, since the soil is modelled as a linear elastic material.
(3) intermediate cases, where the soil is neither failed nor elastic, but the actual non-linear
hysteretic properties of the soil are modelled.
A typical example of the first category is the Mononobe-Okabe method [Mononobe and
Matsuo, 1929], [Okabe, 1924], with its variants [Nadim and Whitman, 1983], [Richards and Elms,
1979], [Richards et al., 1999], in which a wedge of soil bounded by the wall is assumed to move as
a rigid block, with the same acceleration as the ground.
In the vein of the second group are the methods proposed by Wood [1973], by Scott [1973], by
Veletsos and Younan [1994a, 1994b, 1997, 2000], by Li [1999] and by Ortigosa and Musante
[1991]. In particular, Wood [1973] analysed the dynamic response of a homogeneous linear elastic
soil, trapped between two rigid walls connected to a rigid base, providing an analytically exact
solution. The approximate model proposed by Scott [1973] represents the restraining action of
the backfill by a set of massless, linear horizontal springs. The stiffness of the springs is defined
as the subgrade modulus. Veletsos and Younan [1994a, 1994b, 1997, 2000] improved Scotts
model, by using a set of semi-infinite, elastically supported, horizontal bars with distributed mass,
to include the radiational damping of the soil and by using horizontal linear springs of constant
stiffness, to model the horizontal shearing action of the stratum. Li [1999], following Veletsos
and Younans analyses, introduced the effects of foundation flexibility and damping into the
model. In this study, the rigid wall and the viscoelastic backfill are considered to rest on a
viscoelastic half-space foundation. Ortigosa and Musante [1991] proposed a simplified kinematic
method, in which the wall is supported in several locations and the possible wall movement is the

36
Chapter 5 Dynamic soil pressures on a retaining wall

flexural deformation. The free-field shear modulus is used directly by the authors to calculate the
subgrade modulus, without considering the variation of the stress field near the wall.
The elastic approach is not widely accepted in current design procedures, because early elastic
approaches assumed the wall to be rigid and fixed against both deflection and rotation, such that
the soil pressures and forces so computed are significantly larger (2.5 to 3 times, according to
Veletsos and Younan [1997] and 1 to 2 times according to Li and Aguilar [2000]) than those
calculated using limit-state methods. However, some studies carried out by Veletsos and Younan
[1994a, 1994b, 1997, 2000] have shown that both the magnitude and distribution of the dynamic
wall pressures are very sensitive to base constraints and wall flexibility. Therefore, if realistic
values are used both for wall and base constraints, the computed pressures are significantly lower
than the values obtained with fixed-base rigid walls and potentially of the same order of
magnitude as those computed by the Mononobe-Okabe method. Moreover, Li [1999] showed
that, if foundation effects are taken into account, the computed base shear may be of the same
order of that estimated with Mononobe-Okabe, even for a rigid gravity wall. Therefore, after
these studies, the initial limitations to the elastic approach seem to be overcome and this method
can be considered a valuable tool for the seismic design of non-yielding walls.
Finally, an example of the third group of methods is the work by Richards et al. [1999], in
which consideration is given to the plastic response characteristics of the soil in the free-field,
where horizontal stress increases nonlinearly beyond yield. The soil is modelled by a series of
springs, the stiffness of which is given by the subgrade modulus, determined by knowing the
value of elastic or secant shear modulus of the soil in the so-called free-field. Again to the third
group belongs the work of Siller et al. [1991], on the behaviour of gravity and anchored walls.
Three of the methods cited above (one limit state and two elastic approaches) have been
implemented in the current work. They will be described in more details in the following
sections.

5.2 Method by Mononobe and Okabe

5.2.1 Description of the model


The Mononobe-Okabe (M-O) method [Mononobe and Matsuo, 1929], [Okabe, 1924] is a direct
extension of the static Coulombs theory to pseudo-static conditions: pseudo-static vertical and
horizontal accelerations are applied to a Coulomb wedge and the pseudo-static soil thrust is then
obtained through equilibrium of forces acting on the wedge (see Fig. 5.1). The original method

37
Chapter 5 Dynamic soil pressures on a retaining wall

has been developed only for the case of active earth pressures; later, it has been extended
(without any experimental validation) to the case of passive pressures.
The main assumptions involved in the method are:
The wall is assumed to displace sufficiently at the base to mobilise the maximum
shearing resistance of the backfill, so that the classical Coulombs wedge theory can be
applied.
The soil is assumed to satisfy the Mohr-Coulomb failure criterion.
Failure of the soil is assumed to occur along a plane surface, passing through the toe of
the slope and inclined at some angle from the horizontal.
The wedge of soil behind the wall behaves as a rigid body and its acceleration is
uniform.
The wedge of soil bounded by the wall and the failure plane is assumed to be in
equilibrium under the forces acting on it: gravity, earthquake and boundary forces along
the wall and the failure surface (see Fig. 5.1).
The earthquake loading is accounted for by equivalent static horizontal and vertical
forces khW and kvW (with kh and kv horizontal and vertical earthquake coefficients and
W the weight of the wedge of soil), applied at the centre of gravity of the wedge. The
use of constant coefficients implies that dynamic amplification due to site effects
(propagation of seismic waves through a soft layer of soil, resonance effects,
topographic effects, etc.) is neglected. The selection of the appropriate earthquake
coefficients is not a simple matter. Since the peak ground acceleration occurs only for a
very short period during an earthquake, it is clear that kh should correspond to a certain
fraction of the peak ground acceleration; kv will be a fraction of kh. Different
percentages of the maximum acceleration have been suggested in the literature [DAyala
and Speranza, 2002], but consensus has not yet been reached. It can be noted that EC8
[2003] has proposed a relation between allowable wall displacement and percentage of
maximum acceleration to be used in the M-O method.

38
Chapter 5 Dynamic soil pressures on a retaining wall

Fig. 5.1: forces acting on an active wedge in M-O analysis, for a dry cohesionless backfill (from Wood
[1973])

A positive horizontal seismic coefficient causes the total active thrust to exceed the static active
thrust whilst it causes the total passive thrust to be less than the static passive thrust. Therefore,
the active case has been considered, since it is the more critical and the formulae reported from
now on will refer to this case. It is recalled that the active pressure condition corresponds to a
situation in which the soil is exerting a force on the retaining system, whilst the passive pressure
condition corresponds to the opposite case of the retaining system exerting a force on the soil.
The total active thrust can be obtained as:

1 (5.1)
PAE = K AE h 2 (1 k v )
2

where is the unit weight of the soil, H is the height of the wall, kv is the vertical earthquake
coefficient, defined as kv = av / g, with av the vertical pseudo-static acceleration. Finally, KAE is
given by:

cos 2 ( )
K AE = 2
sin( + ) sin( i ) (5.2)
cos cos cos( + + ) 1 +
2

cos( + + ) cos(i )

39
Chapter 5 Dynamic soil pressures on a retaining wall

in which, for a dry soil, the angle is calculated as:

kh
= tan 1 (5.3)
1 kv

The meaning of the different angles appearing in the Eq. (5.2) is explained graphically in Fig.
5.1.
The method, as described above, gives the magnitude of the total force acting on the wall, but
does not state anything about its point of application or the pressure distribution. Although the
M-O method seems to imply that the total thrust acts at a height of h/3 above the base of the
wall, experimental results [Seed and Whitman, 1970] suggest that, under dynamic conditions, it
actually acts at a higher point. The total active thrust PAE, determined from Eq. (5.1) can be
subdivided into a static component Rst, acting at h/3 from the base and a dynamic component
PAE, acting at approximately 0.6h [Seed and Whitman, 1970]. The static component can be
obtained using the classic Coulombs theory (see Eqs. (4.1) and (4.2)), while the dynamic
component can be easily obtained from the difference between the total and the static parts. On
this basis, the total active thrust will act at a height

Rst (h / 3) + PAE (0.6h)


H= (5.4)
PAE

The formulae reported above refer to the case of a dry cohesionless backfill. The method can
easily be extended to include soil cohesion, by considering the equilibrium of the wedge with the
addition of cohesive forces acting along the wall boundary and the failure surface.
The M-O method, very simple and straightforward, has been widely used by designers, because
experimental and theoretical studies have shown that it gives satisfactory results in cases where
the backfill deforms plastically and the wall movement is large and irreversible [Whitman, 1990].
However, there are many practical cases, such as massive gravity walls or basement walls braced
at top and bottom, where the wall movement is not sufficient to induce a limit state in the soil,
and the dynamic soil pressures should be computed using elastic or viscoelastic methods, such as
those explained in more detail in the following sections.

40
Chapter 5 Dynamic soil pressures on a retaining wall

5.2.2 Shortcomings of the model


1. Representation of the complex, transient dynamic effects of earthquake shaking by a
single constant unidirectional pseudo-static acceleration is a crude approximation. The
vertical and horizontal pseudo-accelerations are chosen to be significantly less than the
peak ground accelerations expected to occur during the design earthquake. This reduction
seems to be based on the acceptance of a certain permanent outward movement of the
wall. The amount of reduction to be applied is not clearly defined, even if EC8 [2003]
suggests some criteria for its evaluation.
2. No accounting for the flexibility of the wall is possible in the analysis.
3. The method is not applicable to soils experiencing a significant strength degradation
during earthquakes, such as potentially liquefiable soils.
4. Resonance effects and amplification of earthquake motion as a result of propagation of
seismic waves through a relatively soft soil layer behind the wall are not taken into
account.

5.3 Method by Scott (1973)

5.3.1 General derivation of the model


Probably the simplest available approximate model for evaluating the dynamic soil pressures
induced by ground shaking on walls retaining an elastic stratum is the one proposed by Scott
[1973].
The soil-wall system considered in this work is a semi-infinite uniform layer of viscoelastic soil,
free at its upper surface, bonded to a non-deformable rigid base and retained along either one or
both of its vertical boundaries by rigid walls. The walls are assumed to be of the same height as
the stratum and their mass is neglected in the model. The walls may be either fixed or restrained
at the base by a rotational spring. Both the walls and the shear beam are presumed to be excited
by the same horizontal ground motion, of acceleration A(t) at any time t. The system considered
is shown in Fig. 5.2.

41
Chapter 5 Dynamic soil pressures on a retaining wall

Rigid wall

x, u
Elastic
k Shear Beam k
, G h

R
b R
L
y Rigid foundation
A(t)

Fig. 5.2: system considered by Scott [1973]

The response of the medium in the absence of the wall (the so-called free-or far-field response)
is evaluated considering it to respond as a base-excited, one-dimensional, vertical cantilever shear-
beam. This beam is attached to the wall or walls by springs, representing the soil-wall interaction.
At each level, y, of the shear beam, a resisting dynamic pressure is developed. This pressure is
proportional to the instantaneous lateral displacement, u, at that level, which is the relative
motion between the shear-beam and the wall at that height. The constant of proportionality is the
spring stiffness K. It is implicit therefore that only normal pressures can develop between the wall
and the soil, that these pressures will be both tensile and compressive and that shearing vertical
stresses are neglected in the model.
In general, the soil properties (density , Youngs modulus E and shear modulus G) vary with
depth, and therefore the density and shearing stiffness of the beam, as well as the spring constant
K, also vary with depth. This variation can be easily accounted for in the model, but for simplicity
the equations reported here refer to the case of soil properties constant with depth.
From the equations of motion of the problem, the eigenfrequencies of the different modes can
be obtained as:

n =
(2n 1)2 2G + 2 K (5.5)
4h 2 b

42
Chapter 5 Dynamic soil pressures on a retaining wall

where n is the number of the mode being considered, h is the height of the walls and b is the
width of the shear beam. The first term on the right hand side of Eq. (5.5) is the frequency of the
unrestrained shear beam, while the second term accounts for the effect of the restraining springs.
For the case of b = L, with L being the distance between the two walls, and assuming the walls
to be rigid, the spring constant K is given by:

8G (1 ) (5.6)
K=
L (1 2 )

The participation factors for each mode, in the case of constant soil properties, can be
calculated as:

cos x n dx
4 sin n h
n = 0
= (5.7)
h
2n h + sin 2n h
cos n x dx
2

where the terms n are obtained from n =


(2n 1)
2h
For a given design earthquake, the maximum deflection at a certain height y for the nth mode is
given by:

S vn n cos(n y )
u mn = (5.8)
n

where Svn is the pseudo relative velocity response at the frequency n, obtained from the design
spectrum, with a reasonable value of damping.
The combination of the contributions of the different modes, performed using the square-root-
of-the-sum-of-the-squares method, gives the deflection at elevation y:

2

S cos(n y )

u= (u ) = vn n (5.9)
2

1
mn
1 n

43
Chapter 5 Dynamic soil pressures on a retaining wall

Therefore, the pressure at this elevation y is given by:

p = K u (5.10)

Analyses performed by Scott [1973] have shown that the pressure distribution is dominated by
the first mode contribution and that higher modes are negligible.
Considering only the first mode, the maximum pressure distribution on the wall is given by:

4 KS v1 y y (5.11)
pm1 = cos = p0 cos
1 2h 2h

and the resultant of the dynamic pressure acting on the wall is

2 (5.12)
pm1 = p0 h

From geometrical considerations on the cosine pressure distribution, the point of application of
the above resultant is located at a distance 2h/ from the base of the wall. Therefore, the
maximum moment per linear meter at the base can be calculated as:

2h 4 (5.13)
M m1 = pm1 = p0 h 2
2

5.3.2 Discussion of some details of the model

a) Boundary effects
The model has been developed for the case of two walls, located at a distance L. The boundary
effects exerted by one wall on the other vary, as obvious, with the distance L. When L becomes
large with respect to the height of the walls, h, the restraining effect of the second wall becomes
insignificant, and the pressures exerted by the soil on the considered wall become equivalent to
the pressures exerted by a backfill of unlimited extent. A limit value of L/h = 10 has been

44
Chapter 5 Dynamic soil pressures on a retaining wall

arbitrarily proposed by Scott, to model an infinite backfill. In this case, the same formulae can be
used, but with a value of L/h = 10.
On the other hand, for small values of L/h (1.0 may be assumed as a limit) the model is not
applicable, due to its intrinsic one-dimensional nature. For L/h < 1, indeed, the two-dimensional
nature of the real motion cannot be neglected.

b) Effect of the wall flexibility


Up until now, the walls have been considered to be rigid. However, in reality, a wall deflects
both by bending and by rotation. The wall flexibility is accounted for in the model by considering
a rigid wall hinged at its base, with a rotational spring of stiffness R (refer to Fig. 5.2). With this
spring at the base, if a moment M is applied to the wall, the angular deflection of the wall is thus
given by M = R .
Bending deflections of the walls are neglected, since their incorporation in the model would be
difficult. The effect of wall flexibility can be easily included in the model only in case of a linear
first mode of vibration. This is the case when soil properties increase with depth.
For a linear first mode, the pressure, given by K times the difference between the soil and wall
displacements, increases linearly with height. This is due to the fact that the lateral displacement
of the wall is linear with height (because the wall is rigid) and the dynamic lateral soil
displacement is also linear (because the first mode is linear). The effect of the base rotation of the
wall can therefore be simply considered as a reduction of the effective soil stiffness K. From
moments considerations, the new soil stiffness K can be calculated as:

1
K'= K
Kh 3 (5.14)
1+
3R

5.3.3 Shortcomings of the model


1. The only damping included in the model is that involved in the shear beam itself. For the
case of an infinite backfill, the model indeed does not consider the radiational damping
capacity of the soil. As a result, when the exciting frequencies are close to the natural
frequencies of the stratum, the model may significantly overestimate the response of the

45
Chapter 5 Dynamic soil pressures on a retaining wall

system. This limitation however does not apply when there are two walls, since in this
case the waves are trapped between the walls and no radiation damping can take place.
2. This soil spring stiffness K (or K) is a constant value, which does not depend on the
characteristics of the ground motion. The uncertainty in the model lies in the choice of
this value K, since forces and pressures are proportional to it. Therefore, any error in the
determination of K is directly reflected in the other response quantities. Moreover, as the
Poissons ratio tends to 0.5, K tends to infinity, as well as forces and pressures. This result
is clearly unrealistic.
3. The model is based on the assumption that the demand due to the ground motion is
resisted by shearing action of the medium in the far-field and by column-like extensional
behaviour in the medium between the far field and the medium. The capacity of the
medium adjacent to the wall to transfer forces vertically by horizontal shearing resistance
is not considered.
4. Its a mono-dimensional method and therefore cannot be applied when the soil has
values of width and height comparable. In these cases, indeed, the bi-dimensionality of
the problem cannot be neglected.

5.4 Method by Veletsos and Younan [1994b]

The following description of the method makes reference essentially to the work proposed by
the authors in the paper indicated as 1994b, since this is the model that has been implemented in
this study. Nevertheless, some references are made to the successive works of the authors and to
the modifications introduced to the original model.

5.4.1 Description of the model


The authors investigated a system consisting in a uniform layer of linear viscoelastic material
that is free at its upper surface, bonded to a non-deformable rigid base and retained along one of
its vertical boundaries by a rigid wall. The wall may be either fixed or elastically constrained
against rotation at its base, by a spring of stiffness R. The wall and the soil stratum have the
same height h. Both the base of the layer and the wall are subjected to a space-invariant,
harmonic, uniform horizontal motion, a(t), at any time t., characterised by a frequency and a
maximum amplitude A. The system considered is shown in Fig. 5.3.

46
Chapter 5 Dynamic soil pressures on a retaining wall

Rigid h
wall

x
R
a(t)

Fig. 5.3: System considered by Veletsos and Younan [1994b]

The main hypotheses involved in the method are:


Constant hysteretic, frequency independent material damping, characterised by a
coefficient , which is the same for both shearing and axial deformations.
Soil movement is only induced by horizontal earthquake excitation in the layer bottom.
Therefore it has been assumed in the model that the vertical normal stress does not
change during vibration.
The model has been developed for vertically propagating shear waves, with the
assumption that horizontal variation of vertical displacements in the soil medium is
negligible.
Only normal pressures develop along the wall-soil interface. The wall can resist both
compressive and tensile stresses, which is equivalent to assume that there is complete
bonding between the wall and the retained medium.
According to Veletsos and Younan [1994b], one of the main problems of Scott method [1973]
is its inability to account for the vertical stresses due to horizontal shearing. Therefore, they
proposed to include this variation of horizontal shearing stresses, which can be expressed as:

xy 1 xy (5.15)
= =
y h

with being the dimensionless distance given by = y/h.


Neglecting the horizontal variation of the vertical displacements,

47
Chapter 5 Dynamic soil pressures on a retaining wall

G u G 2u (5.16)
xy = and =
h h 2 2

where u is the relative horizontal displacement of the medium with respect to the moving base.
If u is expressed by the method of separation of variables as a linear combination of modal
terms, through mathematical manipulation [Veletsos and Younan, 1994b], the nth component of
, denoted as ()n, may be recognized to be:

( ) n = n2u n (5.17)

where un is the nth component of the displacement u and n is the circular natural frequency of
the stratum, considered to respond as a cantilever shear-beam, given by

(2n 1)Vs (5.18)


n =
2h

where n refers to the mode being considered and Vs is the shear wave velocity of the stratum.
Eq. (5.17) represents a force per unit of length, that is identical to the force induced by a
massless linear spring of stiffness kn given by:

(2n 1) G
2

k n = n2 = (5.19)
2 h 2

This corresponds to modelling the shearing action of the medium, for each modal component,
with a set of horizontal linear springs of constant stiffness kn, connected at their lower ends to
the common base, subjected to the prescribed ground acceleration A(t). On the other end, the
medium is modelled by a series of semi-infinitely long, elastically supported horizontal bars with
distributed mass, as shown in Fig. 5.4. In reality, it is sufficient to determine the response of a
single bar for each modal component, since the responses of the other bars are proportional to
sin[(2n-1)/2].

48
Chapter 5 Dynamic soil pressures on a retaining wall

x Elastic bar with distributed mass

kn kn kn

Fig. 5.4: Model of the soil stratum: elastically constrained bar

For a viscoelastic bar, with frequency independent damping , subjected to a harmonic motion
of circular frequency , the impedance or dynamic stiffness Kn can be calculated according to Eq.
(5.20). It should be pointed out that is the loss coefficient, equal to twice the damping ratio ,
commonly used in soil dynamics.

K n = ( K st ) n (1 + i )(1 n2 + i ) (5.20)

where n is the dimensionless frequency ratio n = /n and (Kst)n is the static frequency,
defined as

(2n 1) 0 G 2 (5.21)
( K st ) n = with 0 =
2 h 1

The real part of the impedance Kn is related to the force component which is in phase with the
excitation, while the imaginary part represents the component 90 out of phase.
The instantaneous value of the far field, horizontal, steady-state displacement, relative to the
moving base, of a point of the stratum situated at a distance from the base, can be expressed
as the superposition of modal components as [Veletsos and Younan, 1994a]:


(2n 1) it
u f ( , t ) = U n sin e where: (5.22)
n =1 2

16 Ah 2 1 1 (5.23)
Un =
G (2n 1) 1 n2 + i
3 3

49
Chapter 5 Dynamic soil pressures on a retaining wall

If we consider the wall to be elastically restrained against rotation at the base by a spring of
stiffness R, the instantaneous value of steady-state horizontal displacement of the wall at a
distance from the base, is given by:

w( , t ) = h e it (5.24)

where is the complex-valued amplitude of the wall rotation, which can be determined as
explained later.
The above equation, expanded in terms of the natural modes of vibration of the stratum,
becomes:


(2n 1) it
w( , t ) = Wn sin e where: (5.25)
n =1 2

8 (1) n1 (5.26)
Wn = h
2 (2n 1) 2

The instantaneous pressure exerted from the backfill on the wall can then be obtained as the
product of each displacement component and the corresponding bar impedance:


(2n 1) it
( , t ) = K n (U n Wn ) sin e i.e.
n =1 2
8 0
1 1 + i 4
( , t ) = Ah 0 G (5.27)
n =1 ( 2n 1) 1 n + i
2 2 2


(1) n1 (2n 1) it
(1 + i )(1 n2 + i ) sin e
n =1 ( 2n 1) 2

It should pointed out that, in the expression above, the terms involving the displacements Un
represent the pressure induced on a fixed-base wall, while the remaining terms (involving Wn)
represent the contribution of the wall rotation.

50
Chapter 5 Dynamic soil pressures on a retaining wall

The shear and bending moment at the base of the wall can then be easily obtained by
integration of the pressure. It should be pointed out that the last two terms of Eqs. (5.28) and
(5.31) represent the effect of the wall inertia, which may be easily included in this model, with
being the mass per unit of plan area of the wall. The base shear is thus given by:

1
Qb (t ) = Qb0 Qb1 Ah + h 2 2 e it where (5.28)
2

16 0
1 1 + i
Qb0 = Ah 2 (5.29)
3
n =1 ( 2n 1)
3
1 n2 + i

8 0
(1) n1
Qb1 = Gh (1 + i )(1 n2 + i ) (5.30)
2 n =1 ( 2n 1)
2

While the base moment is given by:

1 1
M b (t ) = M b0 M b1 h 2 A + h 3 2 e it where (5.31)
2 3

32 0 (1) n+1

1 + i
M =
0
Ah 3 (5.32)
b
4 n =1 ( 2n 1)
4
1 n2 + i

16 0
1
M b1 = Gh 2 (1 + i )(1 n2 + i ) (5.33)
3
n =1 ( 2 n 1) 3

The rotation amplitude can be obtained from the equilibrium of moments around the wall
base:

1 1
M b (t ) = M b0 M b1 h 2 A + h 3 2 e it = R (5.34)
2 3

It should be pointed out that the response is evaluated here for a harmonic excitation. The
response to an arbitrary ground motion can be determined using Fourier transform techniques.

51
Chapter 5 Dynamic soil pressures on a retaining wall

Once the geometry and the properties of the system have been fixed, the value and distribution
of the pressure acting on the wall are still not determined, because some parameters, such as the
stiffness of the rotational spring and the characteristics of the excitation (frequency and
maximum acceleration for a harmonic motion), still need to be evaluated. The calibration of such
parameters is not trivial and the procedure that has been followed in the present work, in order
to determine these parameters, will be discussed later.

5.4.2 Shortcomings of the model

The rotational stiffness of the spring should be assigned as an input data. In practice,
this stiffness value is difficult to determine a priori and can be known only from results
obtained from a previous dynamic interaction study. In fact, for the elastic behaviour of
a rigid wall, the wall rotation is directly related to the bedrock flexibility. It may be
therefore preferable to express the rotational stiffness by other frequency-dependent
dynamic stiffnesses that are related, in turn, to the soil shear modulus and damping, as
well as to the base geometry of the wall. On the other hand, if the bedrock flexibility is
considered for the wall, it must also be accounted for in the backfill analysis. This can
be done by including the damping capacity of the bedrock along its large contact area
with the backfill, as suggested by Li [1999].
The rotational stiffness is a real-valued quantity. The fact that it does not have a
complex part overamplifies the wall dynamic responses. To overcome this drawback, Li
[1999] replaced this rotational stiffness by the complex-valued dynamic stiffness of a
rigid strip foundation.
In this model, the damping capacity of the wall is neglected. Therefore, all waves
impinging on it cannot be dissipated and the rigid movement of the wall increases the
wave amplitude. As a result, the dynamic wall shear is more amplified for a wall with a
larger rotation, which is not what would be expected.
The assumption of complete bonding between the wall and the retained soil, which
makes possible the development of tensile pressures on the wall, is clearly unrealistic. In
a real case, when the tensile pressures exceed the gravity-induced compressive
pressures, the backfill would tend to separate from the wall and this separation would
increase the wall shears and bending moments with respect to the case of complete
bonding. In fact, with the assumption of complete bonding, the portions of the wall
subjected to tensile pressure will produce contributions to shear and moment of

52
Chapter 5 Dynamic soil pressures on a retaining wall

opposite sign with respect to the compressed zones, thus reducing the total shear and
moment at the base.
The assumption of uniform properties for the soil is also unrealistic, since the shear
modulus of the backfill is likely to increase with depth and this variation will affect both
the magnitude and distribution of the pressure. For a rigid wall, it has been shown
[Veletsos and Younan, 1994b] however that the variation of the soil properties results
in smaller wall forces than in the considered case of constant properties. Therefore it
could be argued that this effect compensates for the effect of the assumption of
complete bonding, which is of opposite sign. However, these two aspects require
further studies.

5.4.3 Further developments of the method [Veletsos and Younan, 1997 and 2000]
As already stated, the method described above is the one proposed by Veletsos and Younan in
1994. More recently, the authors have added some other features to the model. In particular, in
the versions proposed in 1997 and 2000, the flexibility of the wall can also be included in the
model. The effect of wall flexibility is to reduce the forces acting on the wall itself.
Also, different boundary conditions are introduced in the model, with the wall being either free
or hinged at the top. In particular, the force reduction due to wall flexibility is more significant
for cantilever walls than for top-constrained walls, due to the greater effective stiffness of top-
supported walls. However, it was felt that such refinements were not needed within the scope of
the current work.

5.5 Comparison of the results obtained with the three methods implemented
Some comparisons have been made of the results given by the three methods implemented.
First of all, the Veletsos and Younan (V-Y) and Mononobe-Okabe (M-O) methods have been
compared for the case of a massless and fixed-base wall, retaining a backfill of the same height.
The system is subjected to an excitation, characterised by a value of frequency very small with
respect to the fundamental natural frequency of the stratum. This is the so-called static case
[Veletsos and Younan, 1994b], [Li and Aguilar, 2000], which is equivalent to applying the loading
slowly enough not to induce dynamic amplification. Therefore, it is meaningful to compare the
solution obtained with V-Y for the static case to that of M-O, since this latter method neglects
the dynamic amplification due to the backfill or to the wall responses.

53
Chapter 5 Dynamic soil pressures on a retaining wall

The characteristics of the system considered in this example are summarised in Table 5.1. It
should be pointed out that the values of the soil parameters, described in the table, will be kept
constant throughout all this work. In V-Y method, 20 modes have been considered, even if, for
the case of a fixed-base wall, the higher modes are often negligible. In M-O, a horizontal seismic
coefficient equal to the PGA has been used, while the vertical seismic coefficient is equal to half
of the PGA. It has to be noted that, in the M-O approach, only the dynamic contribution is
shown, after the static contribution has been subtracted. The results obtained are shown in Table
5.2.

Table 5.1: characteristics of the system considered

Height [m] Thickness [m] Specific weight [N/m3]


WALL
1 0.4 18000
Mass density [Kg/m3] Poisson ratio [-] Damping coeff.
SOIL
1700 0.3333 0.05
Max acceleration [g] Frequency [rad/s]
EXCITATION
0.25 1

Table 5.2: results obtained from the comparison of Veletsos and Younan and M-O methods, for the
case of fixed massless wall and static excitation
Veletsos et al. M-O
Dynamic base shear [N/m] 3919 1268
Point of application [m] 0.6 0.6
Dynamic base moment [Nm/m] 2346 761

In can be observed that:


the base shear calculated by the elastic approach (3919 N/m) is approximately equal to
AH2 (=4169 N/m), and is approximately 3 times the base shear obtained with M-O
(1268 N/m). This result agrees very well with what reported by Li and Aguilar [2000].
the point of application of the resultant of the pressure obtained with the two methods
is practically identical. This height is close to the 2/ value, corresponding to a pressure
distribution that increases as a quarter-sine from base to top. The latter distribution is
obtained assuming that the soil responds is its fundamental mode of vibration [Veletsos
and Younan, 1997].

54
Chapter 5 Dynamic soil pressures on a retaining wall

From the two previous points, it follows that the base moment obtained with V-Y is
again approximately 3 times that obtained with M-O
The same system has been considered again, subjected to the same static excitation (A = 0.25 g
and = 1), but with the addition of a rotational constraint at the base of the wall in Veletsos
method. The results obtained have been compared to those of the M-O method. It can be
observed that:
as the flexibility increases, the base shear calculated by Veletsos method decreases and
hence approaches the value obtained with M-O; for values of R between 1E6 and 2E7, it
is shown in Fig. 5.5 that the two models give quite close results. This trend was expected
since, as explained in section 5.2, one of the assumptions of M-O method is that the wall
displaces sufficiently at the base to mobilise the maximum shearing resistance of the
backfill. As R decreases and the base of the wall becomes more flexible, the wall
displacements become larger and therefore the results given by V-Y approach those
obtained with M-O.
The variation of the point of application of the resultant of the pressure with R is shown
in Fig. 5.6, for the two models. It can be observed that, as R increases, the result
obtained with V-Y tends to that obtained with M-O, showing a trend which is opposite
to that of the base shear.
On the other hand, the base moment shows a dependence on R which is very similar to
what has been observed regarding the base shear, as shown in Fig. 5.7.

4500

4000 V-Y
3500 M-O
Base shear [N/m]

3000

2500

2000

1500

1000

500

0
1.E+06 2.E+07 4.E+07 6.E+07 8.E+07 1.E+08
R [N/rad]

Fig. 5.5: comparison of base shear, obtained with the two methods, for different values of R

55
Chapter 5 Dynamic soil pressures on a retaining wall

0.7

Point of application of the thrust


0.6

0.5

resultant [m] 0.4

0.3

0.2
V-Y
0.1 M-O
0
1.E+06 1.E+08 2.E+08 3.E+08 4.E+08 5.E+08

R [N/rad]

Fig. 5.6: comparison of the points of application of the thrust resultant, obtained with the two methods,
for different values of R

2500
Base moment [Nm/m]

2000

1500

1000

500 V-Y
M-O
0
1.E+06 1.E+08 2.E+08 3.E+08 4.E+08 5.E+08

R [N/rad]

Fig. 5.7: comparison of base moment, obtained with the two methods, for different values of R

A third comparison between the different methods has been made for the case of the same
system, with a fixed-base wall, excited by a harmonic motion of circular frequency = 400 rad/s
and maximum acceleration A = 0.3 g. In this case, also Scotts method has been included in the
comparison, but only the fundamental mode has been considered, since, according to the author,
higher modes effects are negligible. Concerning M-O method, kh = PGA and kv = kh/2 have
been used. In Veletsos model, the mass of the wall has been neglected and 20 modes have been
considered. The results obtained are summarised in Table 5.3:.

56
Chapter 5 Dynamic soil pressures on a retaining wall

Table 5.3: results obtained with the three method for a fixed-base wall, excited with a frequency of 400
rad/s
Veletsos et al. Scott M-O
Dynamic base shear [N/m] 3647 2328 1701
Point of application [m] 0.63 0.64 0.6
Dynamic base moment [Nm/m] 2306 1482 1021

As expected, the M-O method results in the lowest base shear. This is because this method
relies on the wall movement to relieve the pressure behind the wall [Ostadan and White, 1998].
Also, V-Y method gives the highest base shear. It can be observed, on the other hand, that the
three methods give very close values for the point of application of the resultant of the thrust.

After this quick comparison of the three methods, it has been decided to continue the work using
the model proposed by Veletsos and Younan [1994b] since, as seen above, this method features
the following advantages:
dynamic amplification (induced by the backfill or by the wall) is included in the solution
the capacity of the medium adjacent to the wall to transfer forces vertically by horizontal
shearing resistance is not neglected
the radiational damping capacity of the soil is considered
the wall inertia effect can be included in the model
the flexibility of the wall can be accounted for
Therefore, all the results reported in the subsequent chapters will refer to this method for the
calculation of the dynamic thrust of the soil on the wall.

57
Chapter 6 Combined model computation of capacity

6 COMBINED MODEL COMPUTATION OF CAPACITY

6.1 Introduction
In order to determine the capacity of the considered masonry bridges walls, including also the
effect of the infill material, it has been necessary to combine the model proposed by Veletsos and
Younan [1994b], for the dynamic thrust of the soil on the walls, and the model developed by
Priestley [1985] and Paulay and Priestley [1992], for the seismic response of an isolated wall. Also
the static thrust, calculated according to Coulombs theory, will be considered in the combined
model.
Before going into detail about how this combined model has been realised, one of the
limitations of Veletsos method - the fact that it considers only one wall, thus neglecting the effect
of the presence of a second wall - will be discussed. In particular, it will be shown that, for typical
values of the ratio width of the bridge to height of the walls, this effect is actually negligible.
Then, an overview of the combined model is presented. However, in order to apply this
combined model, some parameters need to be discussed and calibrated. These are essentially the
stiffness of the rotational spring at the base of the wall, the frequency of excitation and the
reduction of the shear modulus of the soil, depending on the level of deformation. Each one of
these aspects will be described in detail in the following sections. Moreover, some other details of
Veletsos method, concerning the maximum acceleration at the base of the wall, will be discussed.
Before going any further, it is important to note that the Veletsos method gives, as results,
complex-valued, time dependent quantities, in which the real part corresponds to the component
which is in phase with the excitation applied to the system, and the imaginary part represents the
component which is 90 out of phase. These quantities vary with time, since the excitation, which
is assumed to be harmonic, is not constant with time. It is evident that, in order to incorporate
this results into the model for the isolated wall, real-valued quantities, constant in time, are
required. Therefore, the modulus of the complex quantities obtained with Veletsos method has
been used. The reason for this choice is that what has actually a physical meaning is the real part
of the response and what is required in the model is the maximum over time of this real part. In
this particular case of a harmonic excitation, characterised by a single frequency and with a given
amplitude, which is constant in time, the maximum of the real part coincides with the modulus of
the complex quantity. This can be easily understood by observing the representation of a given

58
Chapter 6 Combined model computation of capacity

complex quantity, c, as a vector rotating in the imaginary-real plane, with a given modulus, as
shown in Fig. 6.1. It is clear then from the figure that:

max[Re(c )] || c || (6.1)
t

Im

Im(c) c

o Re(c)
Re

||c||
max[Re(c)]

Fig. 6.1: graphical representation of a complex quantity, c

6.2 Parametric study on the reciprocal influence of the two walls


As explained in the previous chapter, it has been decided to calculate the dynamic thrust of the
soil on the wall using the method proposed by Veletsos and Younan [1994b]. Nevertheless, this
method has been developed for a single wall, retaining a uniform layer of viscoelastic soil, whilst
in arch masonry bridges, there are actually two parallel walls, retaining the infill material. It is easy
to understand that, if the two walls are spaced far apart, the pressures on one wall will not be
strongly influenced by the presence of the other. In order to estimate the minimum distance at
which the interaction between the two walls can be neglected and therefore the approximation
introduced by using Veletsos method for the dynamic thrust is acceptable, a parametric study has
been performed, using the model proposed by Wood [1973]. This model considers a
homogeneous linear elastic soil, trapped between two rigid walls, connected to a rigid base, as
shown in Fig. 6.2. The system is subjected to a horizontal, harmonic base acceleration, of
amplitude ah.

59
Chapter 6 Combined model computation of capacity

Fig. 6.2: system considered by Wood [1973]

The resultant of the dynamic thrust (which is the base shear) and the dynamic overturning
moment (about the base of the wall) can be expressed as:

ah
Q = H2 Fp (6.2)
g

ah
M = H3 Fm (6.3)
g

where is the unit weight of the soil, H is the height of the soil, Fm and Fp are the dimensionless
dynamic thrust and moment factors, shown in Fig. 6.3.

Fig. 6.3: dimensionless thrust (left) and moment (right) factors


for various geometries and Poissons ratios [Wood, 1973]

60
Chapter 6 Combined model computation of capacity

The values of dynamic base shear and base moment have been calculated for different values of
the ratio L/H and for the following parameters:
height: H = 2 m
PGA: ah = 0.1 g
Poissons ratio: = 0.33
The results obtained are summarised in Fig. 6.4. It can be noticed that, for L/H > 5 the
variation of base shear and moment becomes approximately nil. This can be even better
appreciated from the observation of the percentage difference between the values obtained with
increasing values of L/H, and the values corresponding to the case of an infinite backfill, taken
as L/H = 8, summarised in Table 6.1. It can be noted that, for L/H = 5, the percentage
difference is less than 1% for both the quantities. Therefore, the ratio L/H = 5 can be viewed as
the limit beyond which the influence of the second wall becomes insignificant and the backfill
can be considered unlimited. Nevertheless, even for L/H = 4, the variation in the base shear is
negligible (1%) and the variation in the base moment (5%) is not very large. Thus, even L/H > 4
can be set as a limit for the unlimited backfill, with an acceptable level of approximation.
Assuming L/H = 4 as the threshold value and considering that typical values of bridge width
range approximately from 4 to 6 m for one rail bridges and from 8 to 12 m for two rails bridges,
it results obvious that Veletsos model is acceptable for values of wall height between 1 and 3 m.
Since, in this work, walls of height ranging from 1 to 2.5 m have been considered, the
approximation introduced by Veletsos method (a single wall with an unlimited backfill) is
applicable.

Influence of the distance between the two walls

710
base shear [N/m] and base

690
moment [Nm/m]

670

650

630

610 base shear


base moment
590

570
2 3 4 5 6 7 8 9
L/H

Fig. 6.4: base shears and base moments obtained for different values of L/H

61
Chapter 6 Combined model computation of capacity

Table 6.1: percentage variation of base shear and base moment


with respect to an infinite backfill, taken as L/H = 8
L/H % variation % variation
base shear base moment
3 12.98 19.4
4 1.08 5
5 0.55 0.99
6 0 0
7 0 0

6.3 Overview of capacity model

6.3.1 Introduction
The combined model consists essentially in the model proposed by Priestley [1985] and Paulay
and Priestley [1992], for the seismic response of an isolated wall, with the addition of the effect of
the infill material. As explained in the previous chapters, the effect of the soil is a thrust, acting
on the wall. This thrust has actually two components: a dynamic component, calculated using
Veletsos and Younan model and a static component, calculated using Coulombs method.

6.3.2 Determination of the acceleration capacity of the wall


The model of the isolated wall, proposed by Priestley [1985] and Paulay and Priestley [1992],
has been described in detail in chapter 3. This model allows to compute the maximum
acceleration that a wall can resist before collapse, through moment equilibrium considerations.
The forces acting on the wall, in absence of infill material, have been shown in Fig. 3.2.
The static soil thrust is included in the model decomposed into a horizontal and a vertical
components, and is applied at h/3. On the other hand, the effect of the dynamic component of
the thrust, calculated according to Veletsos method, is included in the model through a force
RVel, applied at a certain height hVel, as shown in Fig. 6.5. The exact meaning of this force RVel
will be explained in what follows.

62
Chapter 6 Combined model computation of capacity

P
P

RVel

/2
W h
hVel Rst,v
Rst,h
h/2
h/3
o
R

Fig. 6.5: forces acting on the wall in the combined model

It is important to underline that, in the combined model, the wall inertia effects are not
represented anymore explicitly by the distributed force p, as it was in the model for the isolated
wall. This is because, as explained in chapter 5, the wall inertia effect can be easily included in
Veletsos model; in this case, the output given by Veletsos considers both the pressure exerted by
the soil on the wall and the wall inertia. In particular, the resultant of these effects, RVel, is simply
the shear at the base of the wall (multiplied by the gravity acceleration and divided by the
acceleration A), whilst its point of application hVel is a weighted average of the points of
application of the resultants of the two effects, and is calculated from the ratio between base
moment and base shear.
It is important to note that the force RVel can be written in a form proportional to = a/g,
since both the soil pressure and the wall inertia are proportional to the acceleration. This aspect
will be discussed in more detail in a following section.
The acceleration-displacement curve of this combined model (wall plus soil) can be calculated
according to a procedure which is very similar to what has been described in chapter 3 for the
isolated wall. In particular, the seismic coefficient is now given by:

63
Chapter 6 Combined model computation of capacity

W h t
R x P + Rst ,h + Rst ,v (6.4)
= 2 3 2
RVel hVel + P h

where the reaction R is obtained as:

R = P + W + Rst ,v (6.5)

As already explained in section 3.2, the cracking point determination is very important in this
method, for the evaluation of the displacements. The seismic coefficient at cracking, , is
calculated as:

t h
R Rst ,h (6.6)
* = 6 3
RVel hVel + P h

The corresponding displacement at cracking, , is given by:

* P h 3 RVel hVel
2
h 4 Rst ,h 3
* = + h Vel + h (6.7)
EI 3 2 3 81 EI

Eqs. (3.5) through (3.11) described in chapter 3 still hold.


The maximum acceleration that the wall can withstand can thus be obtained as:

amax = max g (6.8)

In order to evaluate the amplification of the seismic input through the wall, which will be
discussed in section 7.3, it is necessary to estimate the first circular frequency of the wall, w..
This frequency can be calculated as:

64
Chapter 6 Combined model computation of capacity

Kn max g (6.9)
w = =
m max

where:
Kn is the secant stiffness corresponding to max on the curve, which can be
evaluated according to:

max g m
Kn = (6.10)
max

max is the displacement corresponding to the attainment of max

It is clear that the value of w just calculated depends on the value of the displacement max.
The method proposed by Priestley [1985] and Paulay and Priestley [1992] is a force-based
procedure, in which the evaluation of the displacement is approximated. Therefore also the value
obtained for w is subjected to this approximation. In any case, it will be shown that the influence
of w on the results is negligible.

6.4 Calibration of the parameters of Veletsos and Younan [1994b] method

6.4.1 Stiffness of the rotational spring


From a physical point of view, the wall does not have a rotational capability at its base, because
it is more or less clamped in the arch (see Fig. 2.1). Nevertheless, a fictitious rotational stiffness
can be used to account for the flexibility of the wall, due to its own deformation. That flexibility
indeed reduces the earth thrust, as the rotational stiffness does [Veletsos and Younan, 1997].
Since the impact of the wall flexibility on the dynamic thrust is not taken into account in the
model developed in this work, the approach of a fictitious rotational stiffness is considered to be
a valuable alternative.
The stiffness of the rotational spring R is a very important parameter in V-Y method. It is to
be noted, indeed, that a variation of the value of R does not only determine a different value of
the resultant of the pressure on the wall, but it also determines a different shape of the pressure
distribution. This effect is very clear for the case of a statically excited system (in the meaning

65
Chapter 6 Combined model computation of capacity

explained in section 5.5), as can be noticed from Fig. 6.6, where the normalized pressure /Ah
is plotted. The results in Fig. 6.6 refer to the massless wall of case 1 (see Table 3.1), subjected to
an excitation of frequency = 1 and maximum acceleration A = 0.2 g. It can be observed that,
as the flexibility at the base of the wall increases, the wall pressures decrease and change their
shape.
1.2

R = 1E20
R = 1E8
1.0
Adimensional height = y / h

R = 1E9 R = 1E7

0.8

0.6

0.4

0.2

0.0
-2 -1 0 1 2
Adimensional pressure / (Ah)

Fig. 6.6: pressure distribution on the wall for different values of rotational stiffness R

The change in shape arises from the higher modes contributions, which become more
significant as the flexibility increases. These results are consistent with what has been observed by
Wood [1973] and by Veletsos and Younan [1994b]. Therefore, a significant number of modes of
vibration is necessary to accurately represent the pressure distribution on walls which are
elastically constrained against rotation at the base. Nevertheless, the base shear is well
approximated using only the contributions of the first two modes of vibration and the base
moment is well approximated using the fundamental mode only. This is shown in Fig. 6.7 and
Fig. 6.8, where the normalised values of these quantities are plotted, for the same system
considered in Fig. 6.6, subjected to an excitation of frequency = 0 and maximum acceleration
A = 0.1 g.

66
Chapter 6 Combined model computation of capacity

1
0.9 2 modes only
0.8 20 modes
0.7 1st mode only

Q b /( Ah )
2
0.6
0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5 6
2
d = Gh /R

Fig. 6.7: effect of wall flexibility on the base shear, for a statically excited system

0.6

0.5 20 modes
1st mode only
0.4
M b /( Ah )
3

0.3

0.2

0.1

0
0 1 2 3 4 5 6
2
d = Gh /R

Fig. 6.8: effect of wall flexibility on the base moment, for a statically excited system

For each one of the walls considered in this study, the stiffness value to be assigned to the
rotational spring at its base has been determined by an iterative process, using a fibre element
software, called Seismostruct [Seismosoft, 2004], developed for nonlinear analyses of structures.
The following steps have been performed:
the dynamic pressure of the soil on the wall is evaluated using Veletsos and Younan
method, with a trial value of R.
A simple finite element model of the wall is constructed. The wall is modelled as a
fixed-base cantilever, made of 3D inelastic beam-column elements. The material

67
Chapter 6 Combined model computation of capacity

constitutive model is elasto-perfectly plastic. The wall is loaded by the pressure


distribution obtained from the previous step.
A static push-over analysis is performed on the model and the moment-curvature plot
at the base of the wall is obtained.
From this plot, it is possible to read the effective rotational stiffness corresponding to
the moment given by Veletsos method.
This stiffness is compared to the initially assumed value and the procedure is repeated
until convergence of the two values is reached.
The values of R obtained for the different walls considered, for the case of = 300, after
application of the procedure described above, are summarised in Fig. 6.9. The case numbers refer
to the different wall typologies and are described in Table 3.1.

Rotational stiffness
1.8E+12

1.5E+12

1.2E+12
R

9E+11

6E+11

3E+11

0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

cases considered

Fig. 6.9: stiffness of the rotational spring for the different cases considered

It can be observed that, as expected, the stiffness of the rotational spring increases as the
dimensions of the walls augment (with increasing case number). Also, it should be pointed out
that, despite some differences in the stiffness values for the various cases, all the walls can be
considered to be fixed-base, since the values of the rotational stiffness are very high. This means
that, in the current work, the flexibility of the wall is neglected and the approach of using a
fictitious rotational spring to account for it would have been a valuable alternative if finite
equivalent values of stiffness were found, which is not the case.

68
Chapter 6 Combined model computation of capacity

6.4.2 Reduction of shear modulus


It is well known that the deformation characteristics of soil are highly nonlinear and this is
manifested in the shear modulus and damping ratio, which vary significantly with the amplitude
of shear strain under cycling loading. In particular, for this work, what is of interest is the
reduction of the shear modulus, for increasing levels of deformation. This effect has been
accounted for in the model, as described in the following.
The low-amplitude shear modulus, or maximum shear modulus, Gmax, which depends essentially
on the characteristics of the soil and on the height of the wall, has been first evaluated according
to the following equation:

m' (6.11)
Gmax = k pa
pa

where:
m' is the mean effective stress
pa is the atmospheric pressure
k is a coefficient, which depends on the relative density. For low values of relative
density, typical of the problems under study, a value of k = 1000 is commonly used
[Pecker, personal communication].

The next step in the evaluation of the reduction of the shear modulus is the selection of a
degradation curve, which is a plot of G/Gmax versus the shear strain . There are many such
curves in the literature, developed for different types of soil.
Since the type of soil used as filling material in masonry arch bridges may vary significantly
from bridge to bridge and this work does not focus on a particular bridge, but is a parametric
study of different possible bridge typologies, it is not an easy task to select a curve. Within this
framework, a curve developed for gravelly soils has been adopted, since it is believed that gravel
has properties which may be considered intermediate among the range of properties of the
different types of soil possibly used as infill material in masonry bridges. In particular, the curve
proposed by Seed et al. [1986] and represented in Fig. 6.10 has been selected.

69
Chapter 6 Combined model computation of capacity

Fig. 6.10: shear modulus reduction curve [Seed et al., 1986]

The last element necessary to evaluate the shear modulus reduction is the level of strain
developed in the soil due to the earthquake, in order to be able to enter the curve and obtain a
value. This shear strain can be easily calculated in Veletsos method, remembering that:

1 u
= (6.12)
h

where u is the displacement, calculated according to Eq. (5.22), and is the dimensionless
height, = y/h. Since u / is a complex quantity, in order to calculate a single value of strain,

the modulus has been used. The quantity expressed by Eq. (6.12), evaluated at mid-height (for
= ), is the maximum shear strain in the soil, max. The reduction of shear modulus has been
evaluated for a level of shear deformation corresponding to 2/3 max. This arbitrary reduction
factor of 2/3 has been introduced in order to be more conservative, since a higher level of
deformation corresponds to a lower value of G/Gmax (as can be seen from Fig. 6.10) and
therefore to a smaller soil thrust.
It is important to outline that, for each case considered, the procedure used to evaluate this
shear modulus reduction is actually iterative, since the shear strain depends on the shear modulus
of the soil which, in turns, depends on the shear strain. Therefore, for each case, it has been
necessary to perform few iterations, until the pair of values of shear strain and modulus belongs
to the selected curve.

70
Chapter 6 Combined model computation of capacity

The values of G/Gmax obtained for the different walls considered, for the case of = 300, after
application of the procedure described above, are summarised in Fig. 6.11. It can be observed
that the amount of reduction of the shear modulus depends on the height of the wall, but not on
its thickness, since both Gmax and do not depend on the thickness of the wall.

Reduction of shear modulus

1.0
0.9
0.8
0.7
G/G max

0.6
0.5
0.4
0.3
0.2
0.1
0.0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

cases considered

Fig. 6.11: reduction of the shear modulus for the different cases considered, for = 300

6.4.3 Frequency of excitation


The results given by Veletsos method are sensitive to the characteristics of the excitation
applied at the base of the wall and, especially, to the circular frequency , which has to be given
as an input. In particular, the height of application of the resultant of the soil thrust does not
depend on the ground motion characteristics, as stated by Veletsos and Younan [1997], but the
value of the resultant is very sensitive to them. Since Veletsos method considers the effect of
dynamic amplification, it is obvious that, if the frequency of excitation approaches the frequency
of the soil layer, the response is amplified, due to the well known resonance effect. This can be
easily demonstrated, by looking at an example. If one considers a system with the characteristics
reported in Table 6.2, subjected to excitations of different frequencies and constant PGA A =
0.1g, the first frequency of the soil layer can be calculated as:

Vs (6.13)
1 =
2h

71
Chapter 6 Combined model computation of capacity

In the case-study analysed herein, 1 results to be 235 rad/s. It is evident, from the results
obtained and reported in Fig. 6.12, that the response (in this case the soil pressure) is significantly
amplified for = 200 rad/s, with respect to the other values of frequency considered.

Table 6.2: characteristics of the system considered

Height [m] Thickness [m] Specific weight [N/m3]


WALL
1 0.4 18000
Mass density [Kg/m3] Poissons ratio [-] Damping coeff.
SOIL
1700 0.333 0.05
Max acceleration [g]
EXCITATION
0.1

1.0 1.0
Adimensional height = y/h

Adimensional height = y/h

0.8 0.8

0.6 0.6

0.4 0.4
=1 =1
= 100 = 100
= 200 = 200
0.2 0.2
= 300 = 300
= 400 = 400
= 500 = 500
0.0 0.0
0 2000 4000 6000 0 2000 4000 6000 8000
3 3
soil pressure [N/m ] soil pressure [N/m ]

Fig. 6.12: variation of the soil pressure distribution for different values of , for a fixed-base wall (left)
and for a rotationally constraint wall, with R = 1.00 E7 (right)

The circular frequency that is applied at the base of the wall in V-Y method has been chosen
as the dominant frequency of oscillation in the transverse direction of the bridge, which is below

72
Chapter 6 Combined model computation of capacity

the walls, because that structure acts as a narrow band filter in the broad band frequency content
of the earthquake. Therefore, the choice of a particular configuration of bridge is equivalent to
the selection of a value of frequency. Few values of frequency have been used in this work, i.e.
= 500, 400, 300, 200 and 100 rad/s. These frequencies correspond respectively to values of the
dominant period of vibration of the bridge in the transverse direction T1 = 0.013, 0.016, 0.020,
0.031 and 0.063. These values have been selected based on the results of a number of 3D finite
element modal analyses. The upper limit of the range of frequencies corresponds to a bridge
characterised by very short and stiff piers. This value of frequency has been successively reduced
to account also for bridges which are less stiff.
When analysing a specific bridge, the value of the dominant frequency of excitation can be
evaluated more accurately. This can be obtained for example running a modal analysis of the
structure, in order to get the periods of vibration. From observation of the response spectrum
corresponding to a given ground motion, in correspondence to the periods of vibration
determined from the modal analysis, it should be possible to identify the dominant frequency of
the bridge, for the chosen input motion. Alternatively, the frequency can also be determined
running a time-history analysis of the bridge, subjected to a given accelerogram, and then
constructing the Fourier spectrum of the response. The dominant frequency is identified by the
peak of the Fourier spectrum.

6.4.4 Discussion on the influence of acceleration on the results


In principle, the capacity of a structure should be independent of the level of seismic excitation
applied. This is true, as expected, also for the described model, but some aspects of it deserve a
more detailed discussion, since they can create some problems.
First of all, in Veletsos method, the maximum acceleration at the base of the wall, A, has to be
given as an input, and its value influences the soil pressure. However, the value of A does not
modify the shape of the soil pressure and hence the height of application of the resultant of the
thrust does not depend on the level of acceleration applied. Moreover, the base shear, which
includes also the effect of the wall inertia, is directly proportional to A. Since the results of
Veletsos method used in the combined model are indeed the point of application and the base
shear normalised with respect to the acceleration (and multiplied by a constant value, g), it is
obvious that the results obtained from this combined model, i.e. the capacity of the different
systems, do not depend on the level of excitation.

73
Chapter 6 Combined model computation of capacity

There are still a couple of aspects to be clarified. As described in the previous sections, the
combined model requires calibration of some parameters, i.e. the stiffness of the rotational spring
at the base of the wall, the reduction of the shear modulus and the frequency of excitation. As
explained previously, the frequency of excitation has been related only to the chosen bridge
typology. This corresponds to the assumption that the range of frequencies considered includes
all the possible frequencies excited for the selected population of bridges, for any given
acceleration. Therefore it can be assumed that the selected values of frequency are independent
of the level of acceleration. On the other hand, the determination of the other two parameters
requires some iterations, which involve other quantities calculated using Veletsos method and
hence apparently dependent on the acceleration. In order to be able to state that the capacity
does not depend on the demand of acceleration, it is necessary to show that also the values of
these two parameters are independent of it.
It can be easily demonstrated that the stiffness of the rotational spring, R, is not influenced by
the value of acceleration. We know from Eq. (5.34) that:

Mb (6.14)
R =

where Mb is given by Eq. (5.31) and can be written, from Eq. (5.34) as:

1
M b0 h 2 A
= 2 (6.15)
1
M b1 h 3 2 + R
3

Since both the terms at the numerator of Eq. (6.15) are directly proportional to A (see Eq.
(5.32) for the term M0b), it can be concluded that also is proportional to A. Similarly, by
looking at Eq. (5.31), it can be easily noted that all the terms defining Mb are directly proportional
to A and hence Mb is proportional to it as well. From this considerations, it is evident, from Eq.
(6.14), that since both the numerator and the denominator are proportional to A, R does not
depend on it.
When using the iterative procedure described in the section 6.4.1, the moment-curvature plot,
obtained from the finite element model, depends only on the shape of the distribution of the soil

74
Chapter 6 Combined model computation of capacity

pressure (which is not influenced by the acceleration) and therefore is independent of A. On the
other hand, the value of dynamic base moment, given by Veletsos method and used to enter the
moment-curvature plot and calculate the corresponding R is proportional to A. However it has
been observed that, since the walls considered are all very stiff, it is correct to use for R the value
of initial stiffness, which is independent of A.
For what concerns the reduction of shear modulus, it is more complicated to demonstrate that
its value does not depend on the level of acceleration. It has been shown that the value of G/Gmax
depends on the value of Gmax and on the level of shear strain, necessary to enter the reduction
curve. Gmax depends only on the geometrical characteristics of the system and on the properties of
the soil. On the other hand, as obvious, the shear strain depends also on the level of acceleration
applied. A parametric study has been performed, in order to investigate the effect of this
dependence of G/Gmax on the acceleration, on the values of capacity, obtained with the combined
model. Only two cases of this parametric study are reported here. The first example concerns the
wall of case 7, subjected to an excitation of frequency = 400 and different levels of
acceleration. The results obtained are shown in Fig. 6.13.

Fig. 6.13: variation of G/Gmax, hVel and max, with the level of acceleration A, for case 7 and = 400

The second example concerns the wall of case 4, subjected to an excitation of frequency =
100. The results obtained for this case are shown in Fig. 6.14.

75
Chapter 6 Combined model computation of capacity

Fig. 6.14: variation of G/Gmax, hVel and max, with the level of acceleration A, for case 4 and = 100

From these two examples it can be observed that the level of acceleration influences the values
of G/Gmax, and this, in turns, influences the point of application of the dynamic soil thrust, hVel
(which on itself does not depend on the acceleration), and the maximum acceleration, max. How
significant is this influence varies with the frequency of excitation. In particular, for the first
example, it can be observed that, even if the values of G/Gmax and of hVel vary significantly, the
results obtained for the capacity (max) are almost constant, with a maximum variation, among the
different cases, of 0.77%. In the second example, the difference in the results for the various
cases is more significant, with a maximum variation of max of 22%.
From the results of this parametric study it can be concluded that:
As A increases, G/Gmax and hence the soil thrust decreases and the capacity of the wall
increases.
For high values of the frequency of excitation , the influence of the level of
acceleration on the reduction of shear modulus and hence on the capacity is negligible.
For lower values of the frequency of excitation , this influence may become
significant.
In what follows, all the results for capacity refer to the case of A = 1 m/s2. For low values of
frequency, in which case the dependence of G/Gmax on the acceleration may be significant, the
results obtained for A = 1 m/s2 represent lower bounds, with respect to the results that would be
obtained with higher levels of acceleration.

76
Chapter 6 Combined model computation of capacity

6.5 Results obtained


The results obtained applying the procedure explained in the previous sections are summarised
in the following figures. In particular, in Fig. 6.15, every single plot refers to a given bridge
typology (identified through the dominant frequency of vibration in the transverse direction) and
represents the maximum acceleration capacity obtained for the different wall configurations
considered.

= 500

0.45
0.4
acceleration capacity [g]

0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

= 400

0.45
0.4
acceleration capacity [g]

0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

cases considered

77
Chapter 6 Combined model computation of capacity

= 300

0.45
0.4

acceleration capacity [g]


0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

cases considered

= 200

0.45
0.4
acceleration capacity [g]

0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

cases considered

= 100

0.45
0.4
acceleration capacity [g]

0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

cases considered

Fig. 6.15: capacity of the walls considered, for 5 different bridge typologies

It is clear from Fig. 6.15 that the maximum acceleration that the walls can survive decreases as
the frequency of excitation decreases, i.e. when the bridge becomes less stiff. In this case, indeed,

78
Chapter 6 Combined model computation of capacity

the frequency of the bridge approaches the value of fundamental frequency 1 of the soil layer
(usually smaller than ), which can be calculated according to Eq. (6.13).
Hence, the soil thrust is amplified, due to the well known phenomenon of resonance. This
effect is particularly significant for values of 0 < < 1.3 1, as evident from Fig. 6.16, where the
dimensionless base shear is plotted versus the ratio of the frequencies /1, for different values
of the wall flexibility d. It should be noted that, as underlined by Veletsos and Younan [1994b],
the amplification factor at resonance increases with increasing flexibility, due to the reduced
capacity of the flexible wall to reflect and dissipate by radiation the waves that are transmitted to
it. It is obvious that, in cases when the soil thrust is amplified due to the resonance effect, the
capacity of the wall decreases.

Fig. 6.16: variation of the base shear with the ratio frequency of excitation - first frequency of the wall,
for different values of wall flexibility d [Veletsos and Younan, 1994b]

The value of the first frequency of the soil stratum, 1, varies from case to case depending
essentially on the height of the soil layer, which is assumed to be equal to the height of the wall,
and on the shear modulus, which, in turns, depends on the frequency of excitation of the system.
Since 1 does not depend on the geometry of the wall, nor on the eventual axial load acting on it,
the different wall cases considered reduce to only 7, characterised by different values of height.
The values of 1 obtained for these cases are summarised in Fig. 6.17. It can be observed that the
frequency of the layer decreases as the height of the soil increases (with increasing case number),

79
Chapter 6 Combined model computation of capacity

as could be expected, since 1 in inversely proportional to h. This can be better appreciated in


Fig. 6.18, where the ratio /1 is plotted versus the height of the soil layer. From this graph, it is
clear that the bridge typology corresponding to a frequency = 100 is the more sensitive to the
resonance effect, since for this case the values of the ratio /1 are close to one.
The frequency of the layer does not change too much from case to case: the minimum value
of 1 encountered in all the cases considered is 1 = 70, whilst the maximum value is 1 = 163.
Also, if we look at any given wall typology, the percentage of variation of 1 with the different
values of ranges between 16% and 39%. Hence, it can be stated that the capacity of the walls is
essentially governed by the frequency of excitation.

180

160

140

120 = 500
1 [rad/s]

100 = 400
= 300
80
= 200
60 = 100
40

20

0
1 5 9 13 17 21 25

cases considered

Fig. 6.17: variation of the first frequency of the soil layer, with the frequency of excitation
and the soil height (= to the wall height)

5 = 500
= 400
4
/ 1

= 300
3 = 200
= 100
2

0
0.5 1 1.5 2 2.5 3

h [m]

Fig. 6.18: variation of the ratio /1 with the height of the soil layer

80
Chapter 6 Combined model computation of capacity

The fact that the capacity of the walls is basically governed by the frequency of excitation can
be better appreciated by looking at some cases of wall configurations more in detail. These are
shown in Fig. 6.19, where the capacity of the cases 1 to 7 is plotted versus the frequency of
excitation . It can be observed that, for all the cases in the figure, the capacity decreases with
decreasing frequency of excitation, i.e. as the bridge typology becomes less stiff. However, the
rate of decrease of the capacity varies from case to case.

0.45
case 1
0.4 case 2
0.35 case 3
case 4
0.3 case 5
a max [g]

0.25 case 6
case 7
0.2
0.15
0.1
0.05
0
100 200 300 400 500
frequency of excitation

Fig. 6.19: capacity of the first 7 cases of wall considered

By comparing the acceleration capacity of two cases of walls of different height, it can be
observed that the higher the wall is, the lower the capacity, for every value of frequency of
excitation . This effect can be appreciated from Fig. 6.20, where the capacity of the wall cases 3
and 5 is plotted. It is recalled that both cases are characterised by a wall thickness t = 0.5 m and
by the absence of an axial load P. The only difference between the two walls is the height, with
case 3 having a height of 1 m and case 5 having a height of 1.2 m.

81
Chapter 6 Combined model computation of capacity

0.45

Acceleration capacity [g]


0.4
Case 3: h = 1 m
0.35
Case 5: h = 1.2 m
0.3
0.25
0.2
0.15
0.1
0.05
0
= 500 = 400 = 300 = 200 = 100
Frequency of excitation [rad/s]

Fig. 6.20: variation of the acceleration capacity with the height of the wall, h

In section 3.3, the maximum acceleration capacity of an isolated wall was calculated, without
considering the soil, which exerts a thrust on it. It is clear from the results reported in this
chapter that, as expected, the presence of the soil has the effect of reducing the acceleration
capacity of the wall. This is true for every value of frequency and for all the wall cases
considered. As an example, the comparison between the capacity obtained for the different cases,
for = 500, with and without considering the effect of the infill material, is reported in Fig. 6.21.

= 500 capacity with soil


0.6 capacity w/o soil
acceleration capacity [g]

0.5

0.4

0.3

0.2

0.1

0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

cases considered

Fig. 6.21: acceleration capacity of the different wall cases obtained with and without considering the
effect of the infill material

82
Chapter 7 Computation of demand

7 COMPUTATION OF DEMAND

7.1 Introduction
Once the capacity of a given wall has been determined, as explained in detail in the previous
chapter, the maximum acceleration that the wall can withstand (amax = gmax) needs to be
compared to the demand of acceleration imposed on that wall by a given earthquake (adem). Since
the capacity is referred to the bridge wall, what needs to be determined is the demand on the wall.
Therefore, for a given ground excitation, with a given peak ground acceleration (PGA), it is
necessary to consider two possible amplification effects, in order to obtain the level of
acceleration demand on the wall. The first one is the amplification or deamplification due to
propagation of the input acceleration from the bedrock to the base of the wall, through the soil
profile and the bridge structure. The second amplification is due to the propagation of the
seismic input through the wall, from its base to, for example, the barycentre.

7.2 Amplification of the seismic input to the base of the wall


Given an earthquake, with a given PGA, applied at the bedrock below a bridge, the acceleration
experienced at the base of the walls, A, may be either amplified or deamplified with respect to
the PGA. The seismic input has indeed to propagate first through the soil profile and then
through the bridge structure.
If the demand is determined using a response spectrum, as it has been done in this work, it is
first necessary to define the spectrum. The shape and amplitude of the spectrum depend on the
characteristics of the soil profile (through the average shear velocity on the upper 30 m), on the
importance of the structure being considered (through the importance factor I, assumed in this
work to be 1) and on the seismic zone the bridge site belongs to (which determines the design
PGA).
Once the acceleration response spectrum has been constructed, the acceleration at the base of
the wall depends only on the dynamic characteristics of the bridge structure, through the
dominant period of vibration of the system in the transverse direction.
In this study, different typologies of bridges have been considered, each one identified by the
value of its dominant period of vibration, as already discussed in a previous section. Also,
different spectra have been constructed, according to the new Italian seismic code [O.P.C.M. n.
3274, 2003]: one spectrum for each one of the three soil categories (A, B C and E, and D) and

83
Chapter 7 Computation of demand

for the PGA corresponding to each seismic zone (zone 1, 2, 3 and 4). This means that, for each
bridge typology, 12 different spectra have been used, in order to determine the acceleration
demand at the base of the wall. It should be pointed out that the spectra considered are elastic
response spectra, characterised by a behaviour factor q = 1. This choice is due to the following
considerations:
First of all, we expect the damage in the walls to occur at the very beginning of the
seismic event. This means that, most likely, the other bridge elements are still not
heavily damaged, but only cracked, and hence may still be in the elastic range of
behaviour.
Even if this is not the case, i.e. if the structural elements of the bridge get damaged
before the walls do, considering an elastic spectrum, instead of reducing it to account
for q, is on the safe side.
Moreover, the ductility of unreinforced masonry bridges, not of very recent
construction, such as most of the Italian masonry arch bridges, is not expected to be
very high and hence the corresponding behaviour factor would be small.
Finally, it is important to remark that the eventual use of a spectrum reduced to
account for q, does not imply any modification in the method proposed, but it would
only produce lower levels of demand.
An example of the spectra obtained, for the different soil categories, for the seismic zone 4
(PGA on rock = 0.05 g) is shown in Fig. 7.1.

Horizontal elastic acceleration response spectra


0.20

0.16
soil cat. A
soil cat. B, C, E
0.12 soil cat. D
Sa [g]

0.08

0.04

0.00
0.0 1.0 2.0 3.0 4.0 5.0

Period [s]

Fig. 7.1: example of elastic horizontal response spectra, for seismic zone 4 (PGA = 0.05 g)

84
Chapter 7 Computation of demand

7.3 Amplification of the seismic input through the wall


The acceleration demand A obtained from a given spectrum, which is the acceleration at the
base of the wall, may be again amplified or deamplified through the wall.
As described previously in this work, the wall can be considered as a system, excited by a
harmonic acceleration history, of amplitude A (determined from the spectrum, as explained in
the previous section) and frequency . For this kind of system, from classical structural dynamics
theory, the acceleration response factor Ra is given by:

2
1
Ra =
w
2
2 2 (7.1)
1 + 2
w w
where:
is the damping ratio, assumed in this work to be 0.05. This relatively low value of
damping has been selected based on the consideration that the walls are expected to be
damaged in the initial phase of the event, while the other elements are still not very
damaged; hence, there is still a low level of energy dissipation and therefore a low value of
equivalent damping is considered appropriate.
w is the natural frequency of the wall, calculated as indicated in section 6.3.2.
The shape of the acceleration response factor, for the assumed value of damping ratio ( = 0.05)
is plotted in Fig. 7.2.

Acceleration response factor


12
10

8
Ra

6
4

0
0 0.5 1 1.5 2 2.5 3
/ w

Fig. 7.2: acceleration response factor for a damping ratio = 0.05

85
Chapter 7 Computation of demand

Once this factor Ra is calculated for each bridge and wall typology, the acceleration demand on
the wall can finally be computed as:

adem = Ra A (7.2)

For the cases considered, this amplification through the wall is not very significant, since the
factor Ra varies between 1.003 and 1.022, with most of the cases close to the lower bound. This is
due to the fact that >> w, because the bridge stiffness is much greater than the wall stiffness.
Therefore, the error in the determination of w, explained in the section 6.3.2, is negligible for
these cases. On the other hand, in cases when is close to w, this error may become significant.
As an example, the results obtained for the acceleration response factor, for the case of =
300, are plotted in Fig. 7.3.

Acceleration response factor


1.025
1.02
1.015
1.01
Ra

1.005
1
0.995
0.99
1 2 3 4 5 6 7 8 9 10 11121314 15161718 19202122 23242526 2728

different cases considered

Fig. 7.3: acceleration response factors for the case of = 300

7.4 Results obtained


The results obtained applying the procedure explained in the previous sections are summarised
in the following figures. In particular, in Fig. 7.4, the variation of acceleration demand with the
soil category is plotted for all the wall configurations and for the different bridge typologies, all
assumed to belong to the seismic zone 1. It is recalled that the case numbers indicated in the
figures refer to the different wall typologies considered, which are described in Table 3.1.

86
Chapter 7 Computation of demand

The variation of acceleration demand with the seismic zones is plotted in Fig. 7.5, for the soil
category B, C and E and for the different bridge and wall typologies. In Fig. 7.6 the attention is
restricted to a single wall type, i.e. case 12, characterised by h = 1.5, t = 0.7 and P/W = 0.5. The
demand corresponding to the three different soil categories and to the four seismic zones is
plotted, for the different values of . Finally, in Fig. 7.7, the results obtained for the wall of case
17, situated in the seismic zone 4, are reported. This wall has the following characteristics: h = 2, t
= 0.8 and P/W = 0. The plot shows the values of acceleration for the different soil categories
and the different values of .
From the observation of all the plots, these general trends can be identified:
the acceleration demand is amplified when moving to softer soils. This is true for all the
frequencies of excitations, but = 100. In this last bridge typology, the soil category B, C
and E gives a higher acceleration than soil D.
The acceleration demand increases with decreasing , since approaches the natural
frequency of the wall, w..
As expected, the acceleration decreases almost proportionally, when increasing the zone
number and therefore the PGA.

87
Chapter 7 Computation of demand

Seismic zone 1, = 500


0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4 \

0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Seismic zone 1, = 400


0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4 \

0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Seismic zone 1, = 300


0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

88
Chapter 7 Computation of demand

Seismic zone 1, = 200


0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Seismic zone 1, = 100


0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

soil A soil B, C, E soil D

Fig. 7.4: variation of the acceleration demand with the soil category, for the different bridge typologies
and the different walls considered

89
Chapter 7 Computation of demand

Soil category B, C and E, = 500

0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Soil category B, C and E, = 400

0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Soil category B, C and E, = 300

0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

90
Chapter 7 Computation of demand

Soil category B, C and E, = 200

0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

Soil category B, C and E, = 100

0.9
0.8
acceleration demand [g]

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

zone 1 zone 2 zone 3 zone 4

Fig. 7.5: variation of the acceleration demand with the seismic zone, for the different bridge typologies
and the different walls considered

91
Chapter 7 Computation of demand

Case 12, = 500 Case 12, = 400

0.7 0.7
acceleration demand [g]

acceleration demand [g]


0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
soil A soil B,C,E soil D soil A soil B,C,E soil D

Case 12, = 300 Case 12, = 200

0.7 0.7
acceleration demand [g]

0.6 0.6

acceleration demand [g]


0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
soil A soil B,C,E soil D soil A soil B,C,E soil D

Case 12, = 100

0.7
0.6
acceleration demand [g]

0.5 zone 1
0.4
zone 2
0.3
0.2
zone 3
0.1
0 zone 4
soil A soil B,C,E soil D

Fig. 7.6: variation of the acceleration demand with the seismic zone and the different soil categories, for
the wall of case 12

92
Chapter 7 Computation of demand

Case 17, seismic zone 1 Case 17, seismic zone 2


0.8 0.6
0.7

acceleration demand [g]


acceleration demand [g]

0.5
0.6
0.5 0.4

0.4 0.3 c

0.3 0.2
0.2
0.1
0.1
0.0 0.0
soil A soil B,C,E soil D soil A soil B,C,E soil D

Case 17, seismic zone 3 Case 17, seismic zone 4


0.35 0.12
acceleration demand [g]

acceleration demand [g]


0.30 0.10
0.25
0.08
0.20
0.06
0.15
0.10 0.04

0.05 0.02
0.00 0.00
soil A soil B,C,E soil D soil A soil B,C,E soil D

= 500 = 400 = 300 = 200 = 100

Fig. 7.7: variation of the acceleration demand with the bridge typology, for the wall of case 17

93
Chapter 8 Comparison of capacity and demand

8 COMPARISON OF CAPACITY AND DEMAND

The results obtained for the acceleration capacity and for the demand in the different cases
considered have been compared. It is obvious that, for the bridge walls to survive, capacity must
be greater than demand. In particular, in most of the figures that follows, attention is restricted to
the ratio D/C between demand and capacity. In order to derive more detailed conclusions, it may
be of some interest not only to check if a wall fails or survives, but also to quantify how far it is
from the limit condition, corresponding to D/C = 1.
In Fig. 8.1, a first comparison between capacity and demand is presented, for all the bridge
typologies () considered and for all the wall cases. This comparison is reported only for the case
of bridges situated in the seismic zone 2 and on soil of type B, C, E, which is the intermediate
type of soil. It can be observed that the demand increases as the frequency decreases, since all the
frequencies considered belong to the first ascending branch of the spectrum (see Fig. 7.1). At the
same time, the capacity generally decreases as the bridge becomes less stiff ( decreases), as
shown in section 6.5. Therefore, as the frequency decreases, the number of walls having
problems increases and, for < 400, almost all the bridge walls situated in seismic zone 2 and on
soil of type B, C, E experience a demand which is greater than their capacity.

94
Chapter 8 Comparison of capacity and demand

= 500, seismic zone 2, soil B,C,E


0.6
0.5
acceleration [g]

0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

= 400, seismic zone 2, soil B,C,E


0.6
0.5
acceleration [g]

0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

= 300, seismic zone 2, soil B,C,E


0.6
0.5
acceleration [g]

0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

95
Chapter 8 Comparison of capacity and demand

= 200, seismic zone 2, soil B,C,E


0.6
0.5
acceleration [g]

0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

= 100, seismic zone 2, soil B,C,E


0.6
0.5
acceleration [g]

0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
cases considered

capacity demand

Fig. 8.1: comparison of acceleration capacity and demand, for all the bridges and walls cases considered,
assumed to be on soil of type B, C, and E and in seismic zone 2.

In order to study more in detail the ratio demand-capacity of the bridge walls considered, this
ratio has been plotted, for each bridge typology and for the total population of bridges
considered, in Fig. 8.2. As explained previously, the ratios demand-capacity have been subdivided
in 4 groups, in order to study more in detail the level of safety of each single case or, on the other
hand, the amount by which demand exceeds capacity. From the different plots, it can be
observed that, as already outlined more than once, the situation gets worse as the frequency of
oscillation of the bridge decreases, i.e. as the bridge becomes less stiff. For 300, the
percentage of cases in which D/C > 1 is almost the same as those with D/C < 1, and the

96
Chapter 8 Comparison of capacity and demand

subdivision among the 4 intervals of D/C is almost uniform. As decreases, the number of walls
experiencing problems increases and, for 200, the number of cases where D is more than 1.5
C is significantly greater than the number of cases falling in each one of the other three intervals.

= 500 = 400 = 300

= 200 = 100 All frequencies

D/C <= 0.5 0.5 < D/C <= 1 1 < D/C <1.5 D/C > 1.5

Fig. 8.2: ratio between demand and capacity obtained for the different bridge typologies

In Fig. 8.3, something similar to Fig. 8.2 is plotted. In this figure, the ratio D/C is studied for
each bridge typology, but restricting the attention each time to the bridges situated on a single
type of soil. In the last 3 plots of the figure, the situation of all the bridges and walls typologies
together is represented, for each soil type. From comparison of these plots, it can be observed
that, for a given bridge typology ( fixed), the number of bridges having problems increases as
the soil becomes softer, as already outlined previously in this work.

97
Chapter 8 Comparison of capacity and demand

= 500, soil A = 500, soil B,C,E = 500, soil D

= 400, soil A = 400, soil B,C,E = 400, soil D

= 300, soil A = 300, soil B,C,E = 300, soil D

= 200, soil A = 200, soil B,C,E = 200, soil D

98
Chapter 8 Comparison of capacity and demand

= 100, soil A = 100, soil B,C,E = 100, soil D

all frequencies, soil A all frequencies, soil B,C,E all frequencies, soil D

D/C <= 0.5 0.5 < D/C <= 1 1 < D/C <1.5 D/C > 1.5

Fig. 8.3: ratio between demand and capacity obtained for the different bridge typologies and for the
different soil types

In Fig. 8.4, the variation of the ratio D/C with the seismic zone is studied for each bridge
typology. In the last 4 plots of the figure, the situation of all the bridges and walls typologies
together is represented, for each seismic zone. As expected, for each given bridge typology (
fixed), the situation becomes critical for an increasing number of bridge walls, as the attention is
moved from the seismic zone 4 towards the other seismic zones, characterised by higher values
of PGA. On the other hand, for a fixed seismic zone, the results confirm what has already been
observed, i.e. the fact that the number of cases subjected to a demand greater than their capacity
increases as the frequency of excitation decreases.

99
Chapter 8 Comparison of capacity and demand

= 500, seismic zone 1 = 500, seismic zone 2 = 500, seismic zone 3

= 500, seismic zone 4 = 400, seismic zone 1 = 400, seismic zone 2

= 400, seismic zone 3 = 400, seismic zone 4 = 300, seismic zone 1

= 300, seismic zone 2 = 300, seismic zone 3 = 300, seismic zone 4

100
Chapter 8 Comparison of capacity and demand

= 200, seismic zone 1 = 200, seismic zone 2 = 200, seismic zone 3

= 200, seismic zone 4 = 100, seismic zone 1 = 100, seismic zone 2

= 100, seismic zone 3 = 100, seismic zone 4 all frequencies, seismic zone 1

all frequencies, seismic zone 2 all frequencies, seismic zone 3 all frequencies, seismic zone 4

D/C <= 0.5 0.5 < D/C <= 1 1 < D/C <1.5 D/C > 1.5

Fig. 8.4: ratio between demand and capacity, for the different bridge typologies and seismic zones

101
Chapter 8 Comparison of capacity and demand

In Fig. 8.5 and in Fig. 8.6, attention is restricted to the wall of case 9, characterised by a height h
= 1.5 m and a thickness t = 0.6 m. The ratio P/W for this case is equal to zero, since no parapet
is considered on top of the walls. The variation of acceleration capacity and demand with the
frequency of excitation is plotted in the two figures.
In Fig. 8.5, each single plot refers to a given seismic zone and the different curves represent the
acceleration demand, for each soil type, and the acceleration capacity (which is only one curve
since it does not depend on the soil category). From observation of the plots, it can be noticed
that in seismic zones 1 and 2, the demand curves are well above the capacity curve, for any value
of frequency and for any soil type. In seismic zone 3, for all the soil types, the walls experience
problems for values of frequency 200. Finally, the walls placed in seismic zone 4 are all safe,
except for the case of = 100, where all the soil categories fail.
On the other hand, in Fig. 8.6, each single plot refers to a given soil type and the different
curves represent the acceleration demand for each seismic zone and the acceleration capacity,
which is again only one curve. It can be observed that, for all soil types, the walls located in
seismic zone 1 and 2 fail for any value of frequency, the walls situated in zone 3 experience
problems for 200, whilst the walls in zone 4 fail only for = 100.

Seismic zone 1
0.8
0.7
acceleration [g]

0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600

frequency [rad/s]

102
Chapter 8 Comparison of capacity and demand

Seismic zone 2
0.8
0.7

acceleration [g]
0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600

frequency [rad/s]

Seismic zone 3
0.8
0.7
acceleration [g]

0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600

frequency [rad/s]

Seismic zone 4
0.8
0.7
acceleration [g]

0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600

frequency [rad/s]

capacity demand, soil A demand, soil B demand, soil D

Fig. 8.5: variation of acceleration demand and capacity with bridge typology, for the wall of case 9 and
for each seismic zone

103
Chapter 8 Comparison of capacity and demand

Soil type A
0.8
0.7

acceleration [g]
0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600

frequency [rad/s]

Soil type B, C, E
0.8
0.7
acceleration [g]

0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600

frequency [rad/s]

Soil type D
0.8
0.7
acceleration [g]

0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600

frequency [rad/s]

capacity demand, zone 1 demand, zone 2 demand, zone 3 demand, zone 4

Fig. 8.6: variation of acceleration demand and capacity with bridge typology, for the wall of case 9 and
for each soil type

104
Conclusions and future developments

9 CONCLUSIONS AND FUTURE DEVELOPMENTS

Even if there are not many real cases of significant damage to masonry arch bridges during
earthquakes, it is believed that these structures may be potentially very vulnerable. In particular,
attention was given to the problem of the interaction between the infill material and the side walls
of the bridges under seismic excitation, in relation with a possible out-of-plane collapse of the
walls.
A capacity model was proposed for the determination of the acceleration-displacement curve of
each bridge wall, including also the thrust exerted on the wall by the infill material. Both the static
and the dynamic components of the soil thrust were evaluated, using appropriate methods. The
filling material has the effect of reducing the capacity of the wall, with respect to the case of an
isolated wall, without soil, presented earlier in the current study.
Moreover, a procedure for the determination of the level of acceleration demand on a bridge
wall was developed, taking into account both the amplification of the acceleration through the
soil layers below the bridge and the bridge structure itself and the amplification through the wall.
This demand was calculated for a wide range of situations, related to the various possible soil
conditions and to the different seismic zones, identified by the new Italian seismic code, which
determine the level of the design peak ground acceleration on rock.
A parametric study on a set of bridges and walls typologies was performed, both for the
capacity and the demand model. For each case considered, acceleration demand and capacity
were calculated and compared between each other, in order to draw conclusions about the level
of safety of that given case.
The proposed procedure is very general, since it allows to calculate the vulnerability of any
given wall, for any level of acceleration. Therefore the method has a very wide application range.
Nevertheless, in the current work, the model was applied only to Italian bridges, since the
response spectra, used for the determination of the acceleration demand, are those defined by the
new Italian seismic code. Therefore, the results obtained are valid only for Italian bridges.

Based on the results obtained in the present work, a number of future developments are
suggested. Firstly, the parametric study proposed in the current study should be expanded,
considering different values of some parameters, such as for example the elastic modulus and the
resistance of masonry. Also, a wider range of cases should be analysed.

105
Conclusions and future developments

The proposed model could be validated against observations of real cases and, possibly,
through comparison with experimental results.
Finally, a finite element analysis of the whole bridge would be very useful to verify the results
obtained and to check the assumptions involved in the model developed in this work. A 3D finite
element model would be the most accurate, but even a 2D analysis would already be very useful
and not beyond the capabilities of most softwares, provided the soil is treated as an elastic
material. It would not be necessary to include a formulation of the Veletsos and Younan method,
since it will be automatically accounted for by the finite element model, if the infill material is
included and provided the soil remains elastic. On the other hand, a finite element model, using a
software which is able to accurately model also the nonlinear behaviour of the soil, would be
necessary, in order to obtain more accurate results, but this is not an easy task.
In conclusion, it is believed that the simplified model proposed here is a good alternative to a
much more complex type of analysis. As already stated, some additional tests on the model could
be useful to validate it.

106
References

10 REFERENCES

ABK [1984]. Methodology for the mitigation of seismic hazards in existing unreinforced
masonry buildings: The methodology, ABK Topical Report 08, El Segundo, California.
Albenga G. [1953]. I ponti: la pratica. UTET Ed., Torino.
ASC Amateur Seismic Centre. Available from URL: http://asc-india.org
DAyala D. and Speranza E. [2002]. An integrated procedure for the assessment of seismic
vulnerability of historic buildings, Proceedings of the 12th European Conference on Earthquake
Engineering, London, Elsevier Science, paper No. 561 (CD ROM).
De Felice G. and Giannini R. [2001]. Out-of-plane seismic resistance of masonry walls,
Journal of Earthquake Engineering Vol. 5, No. 2, pp. 253-271.
Doherty K. T. [2000]. An investigation of the weak links in the seismic load path of
unreinforced masonry buildings, Ph.D. Thesis, Faculty of Engineering of the University
of Adelaide, Department of Civil and Environmental Engineering.
EC8 [2003]. Eurocode 8: Design of structures for earthquake resistance, final draft prEN
1998-5
Galasco A., Lagomarsino S., Penna A., Resemini S. [2004]. Non-linear seismic analysis of
masonry structures, Proceedings of the 13th WCEE Vancouver (in press).
Gambarotta L., Lagomarsino S., Brencich A., De Francesco U. and Resemini S. [2001]. Per
lo studio metodologico e software della capacit portante dei ponti ad arco in muratura.
Parte 1: Inquadramento tipologico per la modellazione geometrica e del materiale.
Gambarotta L., Lagomarsino S., Brencich A., De Francesco U. and Resemini S. [2001b]. Per
lo studio metodologico e software della capacit portante dei ponti ad arco in muratura.
Parte 2: Direttive Tecniche per la verifica di ponti ferroviari ad arco in muratura.
Gambarotta L., Lagomarsino S., Brencich A., De Francesco U. and Resemini S. [2001c]. Per
lo studio metodologico e software della capacit portante dei ponti ad arco in muratura.
Parte 3: Capitolato tecnico per le indagini in sito, necessarie al collaudo di ponti ad arco in
muratura.
Gambarotta L., Lagomarsino S., Brencich A., De Francesco U. and Resemini S. [2001d]. Per
lo studio metodologico e software della capacit portante dei ponti ad arco in muratura.
Parte 4: Analisi tridimensionale ad elementi finiti del ponte per la verifica in esercizio.

107
References

Gisdevelopment [2001]. Seismically deficient structures: engineering lessons, available from


URL: http:// www.gisdevelopment.net
Griffith M. C., Magenes, Melis G. and Picchi L. [2003]. Evaluation of out-of-plane stability
of unreinforced masonry walls subjected to seismic excitation, Journal of Earthquake
Engineering, Vol. 7, Special Issue 1, pp. 141-169.
Kramer, S.L. [1996]. "Geotechnical earthquake engineering", Prentice Hall Ed.
Lagomarsino S. [1999]. Damage survey of ancient churches: the Umbria-Marche
experience, in Seismic Damage to Masonry Buildings, ed. Bernardini, A. Balkema, Rotterdam,
pp. 81-94.
Lam N. T. K., Wilson J. L. and Hutchinson G. L. [1995]. The seismic response of
unreinforced masonry cantilever walls in low seismicity areas, Bulletin of New Zealand
National Society of Earthquake Engineering, Vol. 28, No. 3.
Li X. [1999]. "Dynamic analysis of rigid walls considering flexible foundations", Journal of
Geotechnical and Geoenvironmental Engineering, Vol. 125, No. 9, pp. 803-806.
Li X. and Aguilar O. [2000]. Elastic earth pressures on rigid walls under earthquake
loading, Journal of Earthquake Engineering, Vol. 4, No. 4, pp. 415-435.
Mononobe N. and Matsuo H. [1929]. "On the determination of earth pressures during
earthquakes", Proceedings, World Engineering Congress, p. 9.
Nadim F. and Whitman R. V. [1983]. Seismically induced movement on retaining walls,
Journal of Geotechnical Engineering, ASCE, Vol. 109 (7), pp. 915-931.
Okabe, S. [1924]. "General theory of earth pressures", Journal of the Japan Society of Civil
Engineering, Vol. 12, No. 1.
Ordinanza del Presidente del Consiglio dei Ministri n. 3274 [2003]. Primi elementi in materia
di criteri generali per la classificazione sismica del territorio nazionale e di normative
tecniche per le costruzioni in zona sismica.
Ortigosa P. and Musante H. [1991]. Seismic earth pressures against structures with
restrained displacements, Proceedings of the 2nd International Conference on Recent Advances in
Geotechnical Earthquake Engineering and Soil Dynamics, pp. 621-628.
Ostadan F. and White W. H. [1998]. Lateral seismic soil pressure an updated approach,
Bechtel National San Francisco, California, Presented as part of the US-Japan SSI Workshop, United
State Geological Survey, Menlo Park, California.
Paulay T. and Priestley M. J. N. [1992]. Seismic Design of Reinforced Concrete and Masonry
Buildings, John Wiley and Sons, Inc.

108
References

Picchi L. [2001]. Risposta sismica per azioni fuori dal piano di pareti murarie, Tesi di
Laurea, Universit degli Studi di Pavia, Facolt di Ingegneria, Dipartimento di Meccanica
Strutturale, Relatore Prof. Guido Magenes
Priestley M. J. N. [1985]. Seismic behaviour of unreinforced masonry walls, Bulletin of the
New Zealand National Society of Earthquake Engineering, Vol. 18, No. 2.
Resemini S. [2003]. Vulnerabilit Sismica dei Ponti Ferroviari ad Arco in Muratura, Tesi di
Dottorato, Universit degli Studi di Genova
Resemini S. and Lagomarsino S. [2004]. Sulla vulnerabilit sismica di ponti ad arco in
muratura, XI Congresso Nazionale Lingegneria sismica in Italia, Genova 25-29 gennaio
2004
Richards J. and Elms D. G. [1979]. Seismic behaviour of gravity retaining walls, Journal of
the Geotechnical Engineering Division, ASCE, Vol. 105 (4), pp. 449-464.
Richards, R., Huang C. and Fishman K. L. [1999]. "Seismic earth pressure on retaining
structures", Journal of Geotechnical and Geoenvironmental Engineering, Vol. 125, No. 9, pp. 771-
778.
Scott, R. F. [1973]. Earthquake-induced earth pressures on retaining walls, Proc., 5th World
Conference on Earthquake Engineering, International Association of Earthquake Engineering, Tokyo,
Japan, II, 1611-1620.
Seed H. B. and Whitman R. V. [1970]. Design of earth retaining structures for dynamic
loads, Proceedings of the Special Conference on Lateral Stresses in the Ground and Design of Earth
Retaining Structures, ASCE, New York, N.Y., pp. 103-147.
Seed H. B., Wong R. T., Idriss I. M. and Tokimatsu K. [1986]. Moduli and damping factors
for dynamic analyses of cohesionless soils, Journal of Geotechnical Engineering, ASCE, Vol.
112, No. 11, pp. 1016-1032
Seismosoft [2004]. Seismostruct, a Computer Program for Static and Dynamic nonlinear
analysis of framed structures, available from URL: http//www.seismosoft.com
Siller T. J., Christiano P. P. and Bielak J. [1991]. Seismic response of tied-back retaining
walls, Earthquake Engineering and Structural Dynamics, Vol. 20, pp. 605-620.
Torre C. [2003]. Ponti in Muratura Dizionario Storico-Tecnologico, Alinea Editrice,
Firenze, pp. 394
Veletsos, A. S. and Younan, A. H. [1994a]. Dynamic soil pressures on rigid vertical walls,
Earthquake Engineering and Structural Dynamics, Vol. 23, pp. 275-301.

109
References

Veletsos, A. S. and Younan, A. H. [1994b]. Dynamic modelling and response of soil-wall


systems, Journal of Geotechnical Engineering, Vol. 120, No. 12, pp. 2155-2179.
Veletsos, A. S. and Younan, A. H. [1997]. Dynamic response of cantilever retaining walls,
Journal of Geotechnical and Geoenvironmental Engineering, Vol. 123, No. 2, pp. 161-172.
Veletsos, A. S. and Younan, A. H. [2000]. Dynamic response of flexible retaining walls,
Earthquake Engineering and Structural Dynamics, No. 29, pp. 1815-1844.
Whitman R. V. [1990]. Seismic design and behaviour of gravity retaining walls, Proceedings of
the Special Conference on Design and Construction of Earth Retaining Structures, ASCE, New York,
pp. 817-842.
Wood J. H. [1973]. Earthquake-induced soil pressures on structures, Report EERL 73-05,
Earthquake Engineering Research Laboratory, California Institute of Technology,
Pasadena, California.

110

You might also like