Vous êtes sur la page 1sur 29

Bull Earthquake Eng (2018) 16:397–425

https://doi.org/10.1007/s10518-017-0210-y

ORIGINAL RESEARCH PAPER

Prediction of inter-storey drifts for regular RC structures


with masonry infills based on bare frame modelling

Sanja Hak1,3 • Paolo Morandi1,2 • Guido Magenes1,2

Received: 10 March 2017 / Accepted: 1 August 2017 / Published online: 10 August 2017
 Springer Science+Business Media B.V. 2017

Abstract The traditional construction of masonry infills adjacent to RC structural ele-


ments is still widely adopted in European countries, including seismically active regions.
Given the repeated field observations from damaging earthquakes, pointing to unaccept-
ably high levels of masonry infill damage, the present study is motivated by the need to
improve further the European seismic design approach for new RC structures with masonry
infills, in order to exclude the poor seismic behaviour probably caused by deficiencies in
the verification procedure. Since the in-plane damage to non-structural panels is commonly
controlled through the limitation of inter-storey drifts, the possibility to introduce more
effective verification criteria, accounting for structural properties, infill layouts and
masonry properties is explored. Therefore, starting from the assumption that analyses and
verifications in the design of buildings are commonly accomplished neglecting the pres-
ence of infills, results of extensive nonlinear numerical analyses for different building
configurations are examined. As a result, a simplified procedure for the prediction of
expected inter-storey drifts for infilled structures, based on the corresponding demands of
bare configurations, in function of a simple parameter accounting for structural properties
and the presence of infills, is introduced. Possible implications of the proposed approach
aimed at the improvement of the current design provisions are discussed.

Keywords Seismic design  Masonry infills  RC structures  Infill contribution  Damage


control  Inter-storey drift limitation

& Sanja Hak


sanja.hak@gmail.com
1
European Centre for Training and Research in Earthquake Engineering EUCENTRE, Via Ferrata 1,
27 000 Pavia, Italy
2
Department of Civil Engineering and Architecture, University of Pavia, Via Ferrata 1,
27 000 Pavia, Italy
3
Faculty of Civil Engineering, University of Zagreb, Fra. Andrije Kačića-Miošića 26,
10 000 Zagreb, Croatia

123
398 Bull Earthquake Eng (2018) 16:397–425

1 Introduction

Even though representing a traditional construction technique, masonry infills erected in


full contact with adjacent RC structural elements are still widely used in Europe, including
regions of high seismicity. As part of a numerical and experimental research campaign
related to the seismic response of traditional masonry infills, accomplished at the
University of Pavia in Italy, this work aims to contribute to the development of improved
practical methods for the damage control of masonry infills in the seismic design of RC
structures.
Observations from field surveys after major seismic events repeatedly point towards a
high seismic vulnerability of masonry infills, also in the case of RC structures verified
according to current seismic code regulations. Following some of the damaging earth-
quakes in Europe during recent years, such as in Abruzzo (L’Aquila), central Italy in 2009
(e.g. Braga et al. 2011; Ricci et al. 2011; Vicente et al. 2012), in Lorca, Spain in 2011 (e.g.
Hermanns et al. 2014), in Emilia, northern Italy in 2012 (e.g. Decanini et al. 2012;
Magenes et al. 2012; Ioannou et al. 2012) etc., problems related to the performance of infill
typologies adopted in earthquake-prone European countries have been identified. Reported
levels of damage (e.g. Fig. 1) indicate that losses have to be prevented more efficiently,
either by improvements in the design approach or through the application of alternative
infill typologies.
The seismic response of masonry infilled RC structures has been extensively studied in the
past, based on classical concepts as well as numerical and experimental investigations.
Various aspects have been thoroughly addressed, such as effects on the global response of
newly designed or existing RC structures due to the presence of infills (e.g. Fardis and
Panagiotakos 1997; Magenes and Pampanin 2004), consequences of irregular infill distri-
bution (e.g. Negro and Colombo 1997), detrimental local effects due to increased shear
demands (e.g. Fardis 2006), in-plane failure modes for infills and adjacent structural elements
(e.g. Shing and Mehrabi 2002), out-of-plane response mechanisms (e.g. Flanagan and
Bennett 1999), analytical models for the representation of infills at different levels of
refinement (e.g. Koutromanos et al. 2011; Smyrou et al. 2011; Mosalam and Gunay 2015) etc.
More recently, further important aspects have been recognised and the principal
attention has been shifted towards the concern about acceptable infill damage, including
the influence of simultaneous in-plane and out-of-plane seismic actions. The infill defor-
mation capacity and the attainment of performance criteria assigned to different limit states
has been studied for common masonry typologies, such as slender clay masonry infills with
and without light reinforcement (Calvi and Bolognini 2001; Hak et al. 2013a) and most
recently, reinforced and unreinforced strong clay block masonry infills (da Porto et al.
2013; Morandi et al. 2014). Latest research efforts have also included improvements of the

Fig. 1 Examples of masonry infill damage during the Emilia 2012 earthquake in Italy a, b Moglia;
c Cavezzo

123
Bull Earthquake Eng (2018) 16:397–425 399

infill performance in seismic conditions through the introduction of new products and
innovative construction techniques (e.g. Mohammadi et al. 2011; Preti et al. 2012;
Markulak et al. 2013; Preti et al. 2015).
Regarding the current design approach in line with European seismic code regulations
(CEN 2004a), the limitation of in-plane infill damage is addressed only through the ver-
ification of a single drift limit at the damage limitation limit state, being equal to 0.50% for
the case of rigid masonry infills. Such unique damage control parameter, applied to the
bare structure independently of the structural layout, the infill distribution and the adopted
masonry typology, allows only a rough control of the actual infill damage. Hence, it may
not always be able to guarantee the achievement of efficient design choices and a sufficient
level of safety without imposing unnecessarily oversized RC structural elements. Con-
sidering the response of different RC frame configurations, previous numerical investi-
gations (Hak et al. 2012) have shown that, in terms of in-plane damage, the behaviour of
unreinforced masonry infill walls in RC frames designed in compliance with present
European seismic code regulation can result to be unsatisfactory. Specifically, the current
design approach may be inadequate for sparse and/or weak infills with a low displacement
capacity, while overly conservative requirements may be imposed for densely distributed
and/or innovative infill typologies with a significantly higher displacement capacity.
For practical purposes, multimodal response spectrum analyses are commonly adopted in
the design of buildings and structural models are usually established neglecting the contri-
bution of non-structural elements to strength and stiffness characteristics of the structure.
Hence, masonry infills are considered only in terms of masses and vertical loads, while
analyses and safety verifications are performed on bare frame models. From the practitioners’
viewpoint such choice can be justifiable under certain conditions, considering the strongly
nonlinear response of masonry infills and the complexity of related modelling procedures.
The variety of available infill typologies and construction techniques, as well as the possible
variability of infill layouts during the life of the structure may cause in practise an additional
lack of appropriate attention and expertise at different design stages. Despite such substantial
simplifications in the practical design of RC structures with masonry infills, a satisfactory
seismic response has to be ensured also for the infilled configuration. Accordingly, perfor-
mance requirements defined for damage limitation and life safety limit states have to be
satisfied considering also the masonry infill behaviour.
Motivated by the need to improve current design procedures for the verification of RC
structures with masonry infills, the principal objective of this work is to contribute to a
more effective limitation of inter-storey drifts in the design of new buildings, ensuring
sufficient infill damage control based on the estimation of expected drift demands for
infilled structures. Therefore, an expression that correlates in-plane drifts for bare and
infilled configurations in function of structural and infill properties is proposed. As a result,
the simplified evaluation of inter-storey drift demands for infilled RC structures based on
drift demands for corresponding bare structural models is envisaged.
The major objectives of the study are achieved through the introduction of simplified
parameters used to describe the most relevant properties of considered structural config-
urations and infill properties. The proposed expression is validated based on the inter-
pretation of results obtained from extensive nonlinear dynamic analyses carried out for
prototype structural configurations with different infill typologies. The study addresses
regular RC frame and wall-frame dual structures, newly designed according to European
seismic code regulations with a uniform distribution of masonry infills in plan and
elevation.

123
400 Bull Earthquake Eng (2018) 16:397–425

2 Definition of a density-stiffness coefficient for infilled RC structures

For the correlation of inter-storey drift demands of bare and infilled RC structures the
definition of a coefficient that accounts for the influence of masonry infill strength and
stiffness on the structural response, depending on the relative stiffness of the surrounding
RC structure, is considered appropriate. The structural layouts under consideration are
assumed to be regular in plan and elevation and not significantly affected by contributions
from modes of vibration higher than the fundamental mode in each principal direction.
Hence, torsional effects do not influence the response significantly and have been
neglected. Therefore, the stiffness of the load-bearing system is evaluated independently
for two orthogonal directions.
In line with widely accepted numerical modelling assumptions, in particular when the
global structural response is of primary interest (e.g. Asteris et al. 2011a), masonry infills
can be represented using simple single-strut models of appropriate geometrical properties,
described by a suitable nonlinear stress–strain relationship (e.g. Crisafulli 1997). The
force–displacement response envelope of the infill typically includes several characteristic
points, which can be assigned to different performance limit states and represented by
corresponding strain parameters. Assuming that any value of strain reached in the infill
strut of a single-storey single-bay infilled frame through simple geometric relations can be
related to a corresponding value of inter-storey drift (Hak et al. 2013a), the attainment of
damage limitation and ultimate limit state conditions can be described either in terms of
strain (e0m and eu, respectively) or in terms of inter-storey drift (d0m ¼ dm0 =h and du = du/h,
respectively, where dm0 and du are the related values of horizontal displacement and h is the
storey height). According to the performance objectives in Eurocode 8-Part 1 (CEN
2004a), for masonry infills the limitation of damage and the prevention of failure have to
be ensured. Therefore, inter-storey drifts assigned to limit state conditions with reference to
the infill performance represent significant properties required to characterise the expected
response of a specific infill typology.

2.1 Simplified evaluation of structural stiffness

Values of elastic stiffness assigned to each storey of a RC structure can be approximated


based on the elastic analysis of a bare frame structural model using the equivalent static
force approach. According to Eq. (1) the average elastic structural stiffness KS,j for storey j
is estimated as the ratio of the horizontal storey shear Vj evaluated from Eq. (2), where ns is
the number of storeys, and the inter-storey displacement dr,j given in Eq. (3), where ds,j and
ds,j-1 are average lateral displacements at the top and bottom of storey j, respectively.
Vj
KS;j ¼ ð1Þ
dr;j

X
ns
Vj ¼ Fi ð2Þ
i¼j

dr;j ¼ ds;j  ds;j1 ð3Þ

For the evaluation of the elastic structural stiffness KS,j according to Eq. (1), the dis-
tribution of horizontal forces in elevation is needed. In this study the fundamental mode
shapes and periods in the horizontal directions have been evaluated based on a modal

123
Bull Earthquake Eng (2018) 16:397–425 401

analysis of the structure and the base shear force Fb participating in the first mode has been
found according to Eq. (4), where Se(T1) is the spectral ordinate of the elastic response
spectrum at the fundamental period of vibration T1 and M1 is the corresponding partici-
pating mass. The force Fj acting on storey j has been determined according to Eq. (5),
where sj is the displacement of mass j in the fundamental mode shape, while mj is the
corresponding storey mass associated with gravity loads assumed in the seismic load
combination.
Fb ¼ Se ðT1 Þ  M1 ð4Þ
sj  mj
Fj ¼ Fb  Pns ð5Þ
i¼1 si  mi

Clearly, other simplified procedures may equally be adopted to evaluate the storey
forces Fj; e.g., the simplified lateral force method proposed in Eurocode 8-Part 1 (CEN
2004a).

2.2 Simplified evaluation of masonry infill strength and stiffness

The stiffness of a single masonry infill within a RC structure, referring to bay i in storey j,
may be approximated by the secant stiffness KI,i,j, corresponding to the horizontal force
Fw,i,j, reached at the displacement equal to dm,i,j, as given in Eq. (6a) and illustrated in
Fig. 2a. Analogously, as given in Eq. (6b), where hj denotes the storey height, the secant
stiffness KI,i,j can be expressed in function of the inter-storey drift d0m;i;j that approximately
corresponds to the associated strain e0m;i;j . Considering one load-bearing direction of a
frame or wall-frame dual structural system, the total stiffness KI,j of all infills in storey
j (j = 1…ns, where ns is the number of storeys) can be evaluated as the sum resulting from
the contribution of each single masonry infill, as given in Eq. (7 a,b), where nb,j is the
number of bays in storey j. If for all infills in the same storey masonry typologies with
equal deformation capacities (d0m;i;j ¼ d0m;j ; i ¼ 1. . .nb;j ) are foreseen, the stiffness KI,j
may be expressed by Eq. (7c). On the other hand, if in the same storey the use of infill
types with varying properties in terms of strength (Fw,i,j) and deformation capacity (d0m;i;j ) is
anticipated, a single value of d0m;j may be evaluated according to Eq. (8). In this case d0m;j
represents the deformation capacity of the storey, defined based on the condition that the
same stiffness KI,j is obtained from Eq. (7b) and (7c).

Fig. 2 a Masonry infill secant stiffness; b equivalent diagonal strut model; c estimation of infill strength
reduction due to opening

123
402 Bull Earthquake Eng (2018) 16:397–425

Fw;i;j Fw;i;j
KI;i;j ¼ KI;i;j ¼ ð6a; bÞ
0
dm;i;j d0m;i;j hj
nb;j
X n n
1X b;j
Fw;i;j 1 X b;j

KI;j ¼ KI;i;j ; KI;j ¼ KI;j ¼ Fw;i;j ð7a; b; cÞ


i¼1
hj i¼1 d0m;i;j 0
hj dm;j i¼1
Pnb;j
Fw;i;j
d0m;j ¼ Pi¼1
nb;j Fw;i;j ð8Þ
i¼1 d0m;i;j

To achieve a suitable approximation the horizontal force Fw,i,j may be taken equal to the
peak horizontal strength of the masonry infill calculated according to Eq. (9), where fw,i,j is
the governing masonry strength, tw,i,j the thickness and Lw,i,j the length for infill i in storey
j.
Fw;i;j ¼ fw;i;j tw;i;j Lw;i;j ð9Þ

As discussed e.g. in Hak et al. (2013b), in absence of more specific data, the masonry
strength fw,i,j may be estimated in accordance with Eurocode 8 – Part 1 (CEN 2004a), on
the basis of the corresponding shear strength of bed joints, which may be assumed equal to
the initial shear strength under zero compressive stress fv0,i,j. However, to be more accurate
in the evaluation of the infill strength fw,i,j, more detailed models accounting for different
infill failure modes (e.g. Decanini et al. 1993) should preferably be adopted when possible.
Furthermore, the presence of openings may influence the performance of masonry infills
significantly, as addressed also in a number of recent studies (e.g. Kakaletsis and
Karayannis 2008; Stavridis et al. 2012). Therefore, in the case of realistic building con-
figurations, the possible reduction of strength for perforated masonry infills should be
considered. Assuming the commonly accepted representation of infills by means of
equivalent diagonal strut models (Fig. 2b), a simple reduction factor is typically applied to
decrease the width (or cross section area) of the infill strut to account for the presence of
openings (Fig. 2c), as proposed by several authors (e.g. Dawe and Seah 1988; Al-Chaar
2002; Asteris et al. 2011b; Decanini et al. 2014). The choice of one of these proposals is
considered suitable for a simplified evaluation of the infill contribution and complies with
the procedure proposed in this study.
In order to compare the strength contribution of different infill layouts and masonry
typologies, a reference infill is assumed, representing the case of full infill for masonry
without openings, having a strength equal to fw,0 and a thickness equal to tw,0. Hence, the
relative density parameter wj given in Eq. (10) is proposed, representing a ratio of infill
strength in storey j for any building configuration of interest and the corresponding infill
strength for the specific reference case of full infill.
Pnb;j
fw;i;j tw;i;j Lw;i;j
wj ¼ i¼1 Pnb;j : ð10Þ
fw;0 tw;0 i¼1 Lw;i;j

2.3 Evaluation of average infill versus structural stiffness ratio

Subsequently, for each load-bearing direction of a regular RC structural configuration,


given the distribution of infills and all properties required to characterise the adopted infill
typologies, knowing the infill parameter KI,j and the structural stiffness parameter KS,j for

123
Bull Earthquake Eng (2018) 16:397–425 403

each storey (j = 1…ns, where ns is the number of stories), the relative density-stiffness
parameter Cj can be defined, as given in Eq. (11).
KI;j
Cj ¼ ð11Þ
KS;j

The simplified density-stiffness coefficient Cj represents an average ratio of infill versus


structural stiffness for each storey and allows assessing the combined influence of masonry
infill strength and stiffness with respect to the stiffness of the structural elements on the
building response. A value of Cj equal to zero actually corresponds to a bare structural
configuration, while increasing values of Cj are obtained for stronger infill typologies, due
to higher values of infill strength and/or thickness, as well as for building layouts with more
densely distributed infills. On the contrary, buildings with stiffer load-bearing structural
systems are represented by smaller values of Cj.

3 Comparison of in-plane drifts for bare and infilled RC structures

Based on an extensive numerical study (Morandi et al. 2011; Hak et al. 2012), results of
nonlinear time-history analyses for different structural layouts, infill distributions and
masonry typologies have been evaluated with the aim to assess the correlation of inter-
storey drift demands for bare and infilled RC structural systems in function of different
parameters, which can be accounted for through the proposed density-stiffness coefficient
Cj. The analyses have been carried out considering newly designed two-dimensional RC
frame structures and three-dimensional frame and wall-frame dual systems, meant to
represent characteristic structural configurations, which can be frequently found in the
European building stock. Moreover, three masonry infill typologies have been assumed and
variations of their distribution within the structural system have been accounted for.

3.1 Prototype structural configurations and masonry infill layouts

The selected case studies included two-dimensional RC frames assumed to represent


typical 5.0 m spaced internal structural planes consisting of three bays (5.0, 2.0, and
5.0 m), being part of a simple spatial structural systems with a varying number of stories,
i.e. 3-storey, 6-storey and 9-storey buildings of 3.0 m storey height (Fig. 3).
Each frame configuration was designed following Eurocode design recommendations
(CEN 2002, 2004a, b) for five different levels of seismicity, corresponding to design peak
ground accelerations ag on ground type A equal to 0.05g, 0.10g, 0.15g, 0.25g and
0.35g. Since the buildings were assumed to be founded on ground type B, seismic actions
have been increased by the soil factor S, accounting for the influence of local ground
conditions. The applied soil factor S was equal to 1.2 for the design seismic actions up to
0.15gS, 1.164 for 0.25gS and 1.076 for 0.35gS. Accounting also for the different design
requirements corresponding to two different ductility classes, i.e., medium (DCM) and
high (DCH), 30 different bare frame types have been considered in total. The dimensions
of all structural elements for the two-dimensional frame configurations are summarised in
Table 1.
Furthermore, a group of three-dimensional prototype structures, representing typical
6-storey RC buildings having a storey height of 3.1 m, with spatial frame or wall-frame
dual structural systems (Fig. 4a, b, respectively), has been considered.

123
404 Bull Earthquake Eng (2018) 16:397–425

Fig. 3 Elevation of two-dimensional prototype structures a 3-storey; b 6-storey; c 9-storey

Table 1 Beam and column dimensions of two-dimensional RC frame systems


ag Storey Cross sectional width/height (cm)

3-storey 6-storey 9-storey

DCM DCH DCM DCH DCM DCH

0.05g–0.25g 1st–3rd 35/35 35/35 45/45 45/45 55/55 55/55 External columns
4th–6th – – 35/35 35/35 45/45 45/45
7th–9th – – – – 35/35 35/35
0.35g 1st–3rd 45/45 35/35 50/50 45/45 60/60 55/55
4th–6th – – 40/40 35/35 50/50 45/45
7th–9th – – – – 40/40 35/35
0.05g–0.25g 1st 35/35 35/35 45/45 45/45 55/55 60/60 Internal columns
2nd–3rd 35/35 35/35 45/45 45/45 55/55 55/55
4th–6th – – 35/35 35/35 45/45 45/45
7th–9th – – – – 35/35 35/35
0.35g 1st 45/45 35/35 50/50 45/45 60/60 60/60
2nd–3rd 45/45 35/35 50/50 45/45 60/60 55/55
4th–6th – – 40/40 35/35 50/50 45/45
7th–9th – – – – 40/40 35/35
0.05g–0.35g 1st–3rd 45/24 30/40 50/24 30/45 55/24 30/50 Beams
4th–6th – – 50/24 30/45 55/24 30/50
7th–9th – – – – 55/24 30/50

123
Bull Earthquake Eng (2018) 16:397–425 405

Fig. 4 Plan layouts of three-dimensional case studies a frame system; b wall-frame dual system

Each three-dimensional structural configuration has been designed for two levels of
seismicity, corresponding to design peak ground accelerations ag on ground type A equal
to 0.25 g and 0.35 g. The design has also been accomplished for both ductility classes
(DCM and DCH), resulting in eight different bare configurations, i.e., four frame systems
and four wall-frame dual systems. Dimensions of all structural elements for the three-
dimensional frame and wall-frame dual systems are summarised in Table 2.
According to the building classification in Eurocode 1 (CEN 2002), for the design of all
case studies recommended permanent and variable floor and roof loads have been assigned,
as summarised in Table 3, and common code defined load combinations have been con-
sidered. An estimate of the load contribution due to the presence of infills was included in
the permanent actions and an estimated self-weight of the slab system was accounted for.
Average structural stiffness parameters KS,j for the analysed case studies, evaluated
according to Eq. (1), are summarised in Tables 4 and 5 for the two-dimensional and the
three-dimensional structural configurations, respectively.
In addition to the bare structures, buildings with different infill layouts and masonry
typologies have been analysed, considering three characteristic, commonly adopted types
of traditional unreinforced clay masonry (Fig. 5a).
Typology T1 consists of a 8.0 cm thick single leaf masonry wall constructed of hori-
zontally hollowed brick units with a 1.0 cm thick plaster on each face, typology T2
consists of two 12.0 cm thick leaves, each constructed of horizontally hollowed brick units
divided by an intermediate 5.0 cm cavity and covered with a 1.0 cm thick external plaster,
while typology T3 is constituted by a single 30.0 cm thick leaf built with vertically
hollowed brick units.
Corresponding strength and stiffness properties, assumed based on experimental test
(Calvi and Bolognini 2001), are summarised in Table 6, where Ewh, Ewv represent the
secant modules of elasticity and fwh, fwv the values of compression strength for the hori-
zontal and vertical direction, fwu stands for the shear strength of the bed joints and fws for
the shear strength under diagonal cracking.
For all considered infill typologies, in order to describe the deformation capacity,
characteristic drifts d0m;j equal to 0.30% and du equal to 1.00% (Hak et al. 2012), respec-
tively, corresponding to the attainment of damage limitation and ultimate limit state
conditions, have been assumed. To achieve an illustrative comparison of different con-
sidered infill layouts, the presence of infill T3 in all bays of any structural configuration is
considered to represent a maximum infill density wj = 100%, according to Eq. (10),
selecting the required reference properties accordingly (i.e. fw,0 = 0.30 MPa,
tw,0 = 300 mm). For all infilled structures a constant infill distribution has been assumed in
elevation.

123
406

123
Table 2 Beam, column and wall dimensions of three-dimensional RC frame and wall-frame dual systems
Transversal direction Frame planes Frame planes Wall-frame planes

ag Storey DCM DCH DCM DCH DCM DCH


Cross sectional width/height (cm) Cross sectional width/height (cm)

Frame system Wall-frame dual system

0.25g 1st–3rd 45/45 45/45 External 45/45 45/45 External columns 40/40 45/45 Columns
4th–6th 35/35 35/35 30/30 35/35 30/30 35/35
0.35g 1st–3rd 55/55 45/45 45/45 45/45 40/40 45/45
4th–6th 45/45 35/35 30/30 35/35 30/30 30/35
0.25g 1st–3rd 45/45 45/45 Internal columns 45/45 45/45 Internal columns 30/230 30/230 Walls
4th–6th 35/35 35/35 30/30 35/35 30/230 30/230
0.35g 1st–3rd 55/55 45/45 45/45 45/45 30/230 30/230
4th–6th 45/45 35/35 30/30 35/35 30/230 30/230
0.25g–0.35g 1st–3rd 50/24 30/45 Beams 50/24 30/45 Beams 50/24 30/45 Beams
4th–6th 50/24 30/45 50/24 30/45 50/24 30/45
Bull Earthquake Eng (2018) 16:397–425
Bull Earthquake Eng (2018) 16:397–425 407

Table 3 Floor and roof loads


kN/m2 Floor load Roof load

Permanent actions gk,1 4.0 2.0


Slab system self-weight gk,2 2.5 2.5
Variable actions qk,1 2.0 1.0

Table 4 Structural stiffness parameters for two-dimensional frames


KS,j (kN/m) 3-storey frame 6-storey frame 9-storey frame

ag Storey DCM DCH DCM DCH DCM DCH

0.05g–0.25g 1st 17,880 26,620 27,080 57,370 49,880 121,100


2nd 11,590 21,410 14,430 41,290 25,280 72,610
3rd 9332 19,170 11,980 38,190 20,780 63,430
4th – – 8967 24,760 16,740 47,370
5th – – 7888 22,760 15,120 43,640
6th – – 6398 19,450 13,950 41,040
7th – – – – 10,610 26,650
8th – – – – 9470 24,260
9th – – – – 7114 19,790
0.35g 1st 33,680 26,620 39,920 57,370 60,740 121,100
2nd 18,890 21,410 21,140 41,290 29,680 72,610
3rd 13,230 19,170 17,640 38,190 23,840 63,430
4th – – 20,040 24,760 19,450 47,370
5th – – 18,670 22,760 17,400 43,640
6th – – 14,940 19,450 15,900 41,040
7th – – – – 17,960 26,650
8th – – – – 16,140 24,260
9th – – – – 10,230 19,790

Referring to two-dimensional frame structures, fully (F) and partially (P) infilled con-
figurations, assuming the presence of infill only in the middle bay (Fig. 5b), have been
investigated. Considering three infill typologies (T1, T2, T3) for each structural configu-
ration and infill layout, 180 two-dimensional infilled frames have been analysed in total.
For the case of three-dimensional frame and wall-frame dual structures, nonlinear analyses
have been carried out for the transversal direction, considering three infill distributions of
typology T3, in addition to the bare structures. Specifically, for each structural type, a fully
(F) infilled configuration (wj = 100%) with masonry infills in all bays of the transversal
direction and two intermediate cases, assuming wj & 60% (P2; see Fig. 6a for frame
system) and wj & 30% (P1; see Fig. 6b for wall-frame dual system), have been consid-
ered. Accordingly, 24 three-dimensional structural configurations with different infill
layouts have been analysed.
An overview of all considered structural configurations with different distributions and
types of infill, as well as the corresponding infill density per storey wj calculated from
Eq. (10), is given in Table 7.

123
408 Bull Earthquake Eng (2018) 16:397–425

Table 5 Structural stiffness parameters for three-dimensional frame and wall-frame dual systems along the
transversal direction
KS,j (kN/m) 6-storey frame 6-storey wall-frame dual system

ag Storey DCM DCH DCM DCH

0.25g 1st 163,900 294,000 352,700 568,900


2nd 89,100 198,400 151,400 276,700
3rd 75,700 183,400 107,600 212,700
4th 59,200 120,400 80,600 167,500
5th 55,800 114,800 59,000 132,600
6th 44,500 100,400 48,000 117,800
0.35g 1st 241,100 294,000 352,700 568 900
2nd 118,000 198,400 151,400 276,700
3rd 94,200 183,400 107,600 212,700
4th 78,300 120,400 80,600 167,500
5th 71,100 114,800 59,000 132,600
6th 50,300 100,400 48,000 117,800

Fig. 5 a Infill typologies T1, T2 and T3; b Fully and partially infilled two-dimensional 3-storey prototype
structures

Table 6 Properties of masonry for considered infill typologies


Typology tw (mm) fwh (MPa) fwv (MPa) fwu (MPa) fws (MPa) Ewh (MPa) Ewv (MPa)

T1 100 1.18 2.02 0.44 0.55 991 1873


T2 260 1.11 1.50 0.25 0.31 991 1873
T3 300 1.50 3.51 0.30 0.36 1050 3240

Estimating the total infill stiffness contribution according to Eq. (7a), obtained infill
stiffness parameters KI,j for the two-dimensional and three-dimensional structural config-
urations, fully infilled with typology T3, are summarised in Tables 8 and 9, respectively.
Subsequently, the corresponding infill versus structural stiffness ratios Cj have been
calculated according to Eq. (11), as shown in Tables 10 and 11, respectively, for the two-
dimensional and three-dimensional structural configurations, fully infilled with typology
T3. Larger values of Cj are notably achieved for DCM, in comparison to DCH, as well as

123
Bull Earthquake Eng (2018) 16:397–425 409

Fig. 6 Examples of assumed infill distributions a frame system (P2_T3); b wall-frame dual system (P1_T3)

Table 7 Summary of analyses performed on bare and infilled structures with corresponding infill density wj
wj (%) Bare Partial infill Full infill

B P_T1 P_T2 P_T3 F_T1 F_T2 F_T3


2d frame 0.0 8.2 12.1 16.7 48.9 72.2 100.0
B P1_T3 P2_T3 F_T3
3d frame 0.0 28.5 57.0 100.0
3d wall-frame 0.0 30.1 60.3 100.0

Table 8 Infill stiffness parameters for two-dimensional frames fully infilled with infill T3 (F_T3)
KI,j (kN/m) 3-storey frame 6-storey frame 9-storey frame

ag Storey DCM DCH DCM DCH DCM DCH

0.05g–0.25g 1st 109,500 109,500 106,500 106,500 103,500 102,500


2nd–3rd 109,500 109,500 106,500 106,500 103,500 103,500
4th–6th – – 109,500 109,500 106,500 106,500
7th–9th – – – – 109,500 109,500
0.35g 1st 106,500 109,500 105,000 106,500 102,000 102,500
2nd–3rd 106,500 109,500 105,000 106,500 102,000 103,500
4th–6th – – 108,000 109,500 105,000 106,500
7th–9th – – – – 108,000 109,500

Table 9 Infill stiffness parameters for three-dimensional frame and wall-frame dual systems fully infilled
with infill T3 (F_T3) along the transversal direction
KI,j (kN/m) 6-storey frame 6-storey wall-frame dual system

ag Storey DCM DCH DCM DCH

0.25g 1st–3rd 618,400 618,400 594,200 594,200


4th–6th 635,800 635,800 605,800 605,800
0.35g 1st–3rd 601,000 618,400 594,200 594,200
4th–6th 618,400 635,800 605,800 605,800

123
410 Bull Earthquake Eng (2018) 16:397–425

Table 10 Infill versus structural stiffness ratios for two-dimensional frames fully infilled with infill T3
(F_T3)
Cj [-] 3-storey frame 6-storey frame 9-storey frame

ag Storey DCM DCH DCM DCH DCM DCH

0.05g–0.25g 1st 6.12 4.11 3.93 1.86 2.07 0.85


2nd 9.45 5.11 7.38 2.58 4.09 1.43
3rd 11.73 5.71 8.89 2.79 4.98 1.63
4th – – 12.21 4.42 6.36 2.25
5th – – 13.88 4.81 7.04 2.44
6th – – 17.11 5.63 7.63 2.60
7th – – – – 10.32 4.11
8th – – – – 11.56 4.52
9th – – – – 15.39 5.53
0.35g 1st 3.16 4.11 2.63 1.86 1.68 0.85
2nd 5.64 5.11 4.97 2.58 3.44 1.43
3rd 8.05 5.71 5.95 2.79 4.28 1.63
4th – – 5.39 4.42 5.40 2.25
5th – – 5.78 4.81 6.03 2.44
6th – – 7.23 5.63 6.60 2.60
7th – – – – 6.01 4.11
8th – – – – 6.69 4.51
9th – – – – 10.56 5.53

Table 11 Infill versus structural


Cj [-] 6-storey frame 6-storey wall-frame dual system
stiffness ratios for three-dimen-
sional frame and wall-frame dual ag Storey DCM DCH DCM DCH
systems fully infilled with infill
T3 (F_T3) along the transversal 0.25g 1st 3.77 2.10 1.68 1.04
direction
2nd 6.94 3.12 3.92 2.15
3rd 8.17 3.37 5.52 2.79
4th 10.74 5.28 7.52 3.62
5th 11.39 5.54 10.27 4.57
6th 14.29 6.33 12.62 5.14
0.35g 1st 2.49 2.10 1.68 1.04
2nd 5.09 3.12 3.92 2.15
3rd 6.38 3.37 5.52 2.79
4th 7.89 5.28 7.52 3.62
5th 8.70 5.54 10.27 4.57
6th 12.29 6.33 12.62 5.14

for frame structures in comparison to wall-frame dual systems, reflecting the larger
influence of the infill presence in the case of more flexible structural configurations.
Accordingly, the parameter Cj also increases in elevation.

123
Bull Earthquake Eng (2018) 16:397–425 411

3.2 Summary of assumptions in design and analysis

The parametric numerical investigations relied upon a series of nonlinear dynamic analyses
of the selected case studies. The design for five different levels of seismicity and two
ductility classes was based on results of multimodal dynamic response spectrum analyses,
carried out on linear bare frame models (Morandi et al. 2011; Hak et al. 2012). The design
choices were supplemented with Italian National Code provisions (NTC08 2008), pri-
marily for the definition of site-specific design seismic actions corresponding to selected
sites in Italy for both limit states, i.e., the ultimate and the damage limitation limit state
with reference return periods of TULS equal to 475 and TDSL equal to 50 years, respectively.
Values of the behaviour factor q equal to 3.90 and 5.85 for the frame systems and 2.20 and
3.30 for the wall-frame dual systems have been assumed for DCM and DCH, respectively.
The structural elements were designed complying with capacity design principles, based
on the action effects evaluated from linear analyses. For each ductility class, the specific
provisions for earthquake resistant design related to dimensioning and detailing were
followed and the relevant requirements regarding strength and ductility at the ultimate
limit state were met. Verifications of inter-story drift limitations at the damage limitation
limit state were accomplished for the case of buildings having non-structural elements of
brittle materials attached to the structure, limiting the inter-storey drift demands induced by
the damage limitation seismic action to 0.50%. Possible second-order P–D effects due to
geometric nonlinearities were also included following the code provisions.
To carry out the nonlinear dynamic analyses, evaluating primarily the global response in
terms of inter-storey drift demands, the structural analysis program Ruaumoko (Carr 2007)
has been used and sufficiently simple, yet reliable models have been adopted (Hak et al.
2012). Well-established modelling strategies for structural and non-structural elements
have been followed. Similar approaches have been implemented and validated in previous
studies; see e.g. Magenes and Pampanin (2004) for RC bare frame and infilled frame
configurations or Sullivan et al. (2006) for RC wall-frame structures. Major modelling
assumptions, assumed hysteretic rules and the anticipated frame response mechanism are
illustrated in Fig. 7.
The modelling approach related to RC structural elements has also been validated on
results from in-plane tests of bare and infilled frames (Calvi and Bolognini 2001), as
discussed in Hak et al. (2012). Given the fact that capacity design principles have been
followed a response mechanism with strong columns and weak beams was expected.

Fig. 7 Summary of frame and masonry infill modelling assumptions

123
412 Bull Earthquake Eng (2018) 16:397–425

Each RC member was represented by a one-component model (Giberson 1967) con-


sisting of a frame element, assumed to remain perfectly elastic, with nonlinear rotational
springs representing plastic-hinges at the ends, where all inelastic deformations are con-
centrated. The modified Takeda hysteresis rule (Otani 1981) was assigned to the potential
plastic hinges. The stiffness corresponding to unloading and reloading was characterised
by the coefficients a and b, adopting a = 0.5 for beams, columns and walls; b = 0.3 for
beams and b = 0 for columns and walls. The ratio between initial and post-yielding
stiffness was defined by the Ramberg–Osgood moment–curvature bi-linear factor r,
assuming a value of r = 0.015 for beams and r = 0.005 for columns and walls. No
strength degradation was taken into account and the possibility of shear failure occurrence
was not considered, since appropriate design and detailing according to modern seismic
code provisions was envisaged. Nevertheless, a minimum acceptable representation of
beam-column joint regions (Priestley et al. 2007) was introduced in the frame model to
account for the contribution of beam-column joint flexibility to total lateral displacements
due to strain penetration and joint shear deformations. Therefore, an elastic element of
finite length was assumed in the region between the intersection of beam and column
centrelines and the member end. The connection between RC walls and adjacent beam
elements on each side of the wall was modelled by means of rigid elements having a length
equal to half of the wall length. The influence of a possible shift in neutral-axis depth,
which may occur when the walls deflect up to cracking and subsequent yielding, has been
assumed to be not significant for the response quantities of interest (Sullivan et al. 2006).
Furthermore, a yield moment-axial force interaction diagram was defined for each vertical
frame element, while the yield moments for beams were obtained from a bilinear idealisation
of the moment–curvature relationship, as suggested by Priestley et al. (2007). Accounting in a
simplified manner for the stiffness reduction due to cracking, values of effective initial
stiffness for all structural members were defined as the secant stiffness of the bilinear curve;
for vertical elements the bilinear estimation was based on the moment–curvature relation
corresponding to the axial force occurring in the seismic load combination.
Given the extensive number of analyses and considering the major objectives of the
study, the widely accepted choice to represent masonry infills by means of an equivalent
single-diagonal-strut model has been adopted. This modelling approach is not able to
accurately capture possible unfavourable local effects on RC members during the analysis
due to the interaction between structural frame elements and the masonry infill. However,
since the focus of this work is the evaluation of global inter-storey drift demands, the single
strut model is considered to be adequate. The local effects on the RC elements have been
considered a posteriori, applying a simplified criterion proposed by Hak et al. (2013b).
Moreover, with the aim of achieving a satisfactory representation of selected infill prop-
erties, the infill model has been calibrated based on existing experimental test results (Hak
et al. 2012). The fundamental parameters of the equivalent strut in terms of strength and
stiffness have been defined using the model proposed by Decanini et al. (1993), accounting
for different infill failure modes. Diagonal tension failure has been identified as the gov-
erning mode for all infill typologies.
In order to represent the axial response of the strut model due to cyclic loading, a
masonry strain–stress hysteretic model has been adopted. Despite the fact that several
cyclic models for masonry infills of relatively simple implementation have been developed
more recently, see e.g. Rodrigues et al. (2010), Cavaleri and Di Trapani (2014), based on
detailed model calibrations accomplished in previous studies (Hak et al. 2012, 2013a), the
well-established model by Crisafulli (1997) was found to be suitable for the infill
typologies of interest. The model allows the representation of the cyclic compressive

123
Bull Earthquake Eng (2018) 16:397–425 413

behaviour of masonry by several hysteresis rules, considering different response mecha-


nisms for loading, unloading and reloading; see also Fig. 7. The envelope is independent of
the load pattern and loading history, coinciding approximately with the stress–strain curve
under monotonic loading.
For the definition of the seismic input, consisting of sets of earthquake records com-
patible with the adopted code spectra, previous recommendations (Beyer and Bommer
2007; Iervolino et al. 2008; Masi et al. 2011) have been followed. Consequently, the
application of natural records has been found the most suitable solution. The earthquake
input ground motions have been selected using the computer software REXEL (Iervolino
and Galasso 2009; Iervolino et al. 2010) from the ITalian ACcelerometric Archive
(ITACA; Luzi et al. 2008; Pacor et al. 2011). Two sets of ten records were selected, one
compatible to the elastic response spectrum adopted for design at the ultimate limit state
for the intermediate peak ground acceleration of 0.15gS and one compatible to the cor-
responding elastic response spectrum at the damage limitation limit state for 0.054gS. For
each target spectrum the deviation of the average earthquake spectrum has been kept below
±10% in the range between 0.05 and 3.0 s, while the spectral value at zero period matched
the PGA. Hence, regarding the number of records and the deviation of average earthquake
spectra, more restrictive criteria than required by the provisions for the representation of
the seismic action according to Eurocode 8-Part 1 (CEN 2004a) have been adopted. The
compatible records have been scaled to the remaining PGAs adopted for design, i.e.
0.023gS, 0.040gS, 0.097gS and 0.142gS at the damage limitation and 0.05 gS, 0.10gS,
0.25 gS and 0.35 gS at the ultimate limit state.

3.3 Results of nonlinear time-history analyses

Investigations of the overall nonlinear response of the case studies have shown that plastic
hinges formed at the bottom of the columns in the bottom storey and at the beam ends.
Hence, the expected response mechanism was achieved confirming the satisfactory
application of capacity design principles and the absence of soft-storey mechanisms (Hak
et al. 2012).
The evaluation of response quantities was carried out in terms of average values. In
particular, for the needs of this study, the average of maximum inter-storey drifts attained
in each storey based on analyses for the set of ten earthquake records have been evaluated.
Accordingly, for the different structural configurations and values of infill density wj pairs
of average bare frame drifts dl,j and infilled frame drifts dl,w,j have been identified per
storey. Results obtained for the two-dimensional case study frame configurations at the
damage limitation and the ultimate limit state for both ductility classes, fully infilled with
different infill typologies (T1, T2 and T3), are shown in Figs. 8, 9 and 10, respectively.
Each curve in these plots consists of five points, representing pairs of data obtained for the
same building configuration (i.e. 3-storey, 6-storey or 9-storey frame of ductility class M or
H), designed for five different levels of seismicity (i.e. 0.05 gS, 0.10 gS, 0.15 gS, 0.25 gS
and 0.35 gS) at the corresponding seismic action.
Similarly, the results for three-dimensional frame and wall-frame dual systems are
shown in Figs. 11 and 12, respectively, at the damage limitation and the ultimate limit state
for both ductility classes, with different T3 infill distributions (P1, P2 and F). Each curve in
these plots consists of two points, representing pairs of data obtained for the same building
configuration (i.e. 6-storey frame or wall-frame dual systems of ductility class M or H),
designed for two different levels of seismicity (i.e. 0.25 gS and 0.35 gS) at the corre-
sponding seismic action.

123
414 Bull Earthquake Eng (2018) 16:397–425

Fig. 8 Average drift demands, 3-storey, 6-storey and 9-storey 2d frames (F_T1, wj = 48.9%) a DLS;
b ULS

Fig. 9 Average drift demands, 3-storey, 6-storey and 9-storey 2d frames (F_T2, wj = 72.2%) a DLS;
b ULS

Fig. 10 Average drift demands, 3-storey, 6-storey and 9-storey 2d frames (F_T3, wj = 100.0%) a DLS;
b ULS

Fig. 11 Average drift demands, 6-storey 3d frames (P1_T3, P2_T3 and F_T3) a DLS; b ULS

Fig. 12 Average drift demands, 6-storey 3d wall-frames (P1_T3, P2_T3 & F_T3) a DLS; b ULS

Hence, one curve represents the relation between average drifts of bare and infilled
structures in storey j of a building typology with infill density wj (j = 1…ns), designed for
increasing levels of seismicity in ductility class M or H. The dimensions of RC structural

123
Bull Earthquake Eng (2018) 16:397–425 415

elements of two-dimensional frame configurations and three-dimensional frame and wall-


frame dual systems are the same for different design seismic intensities, resulting in the
same structural stiffness, with exception of the columns in the frame structures designed for
DCM at 0.35 gS. Therefore, representing every infill typology and/or infill layout through
the infill strength and stiffness parameter KI,j, each curve can be associated with one value of
the density-stiffness coefficient Cj. Consequently, the results indicate that for a constant
value of Cj, with increasing bare frame drift demands, the average drift demands of the
infilled frame increase approximately linearly up to a certain level of drift. Changes in the
ratio of average infilled and bare frame drift demands dl,w,j/dl,j B 1.0 are attributed to
changes in either the structural stiffness coefficient KS,j or the infill strength and stiffness
parameter KI,j. After exceeding a certain level of drift, the drift demands of infilled con-
figurations cannot be described in function of bare frame drifts with the same linear relation.
For buildings designed in DCH, having due to the adopted design choices higher values
of initial frame stiffness and higher values of the parameter KS,j than in DCM, drift
demands of the infilled configurations are closer to those of the bare structures, meaning
that the ratio dl,w,j/dl,j is larger (closer to 1.0). Similar observations apply to the com-
parison of three-dimensional frame and wall-frame dual systems; due to a higher initial
stiffness of the bare configurations, the presence of infill has less influence on the response
of the wall-frame dual systems than on the frame systems. On the other hand, for
increasing values of the parameter KI,j, quantifying the presence of infill, values of the ratio
dl,w,j/dl,j are smaller (closer to 0.0), i.e., a more pronounced reduction of drift demands for
the infilled configurations with respect to the bare structures is identified.

4 Procedure for the prediction of in-plane drifts for infilled RC structures

In line with the observations discussed based on the results of nonlinear time-history
analyses carried out for bare and infilled RC structural systems in function of different
building parameters, a relationship is proposed for the evaluation of maximum expected
average inter-storey drift demands for masonry infilled structural configurations, in func-
tion of the average drift demand for corresponding bare structures.

4.1 Estimation of elastic design inter-storey drift demands

In seismic design procedures according to Eurocode 8-Part 1 (CEN 2004a) and the Italian
provisions (NTC08 2008), the evaluation of displacement demands is actually only
required to account for P-D effects and to verify inter-storey drift limits. As specified in the
code, in the case of linear analyses, the displacements at the ultimate limit state ds,ULS
induced by the design seismic action should be calculated on the basis of the elastic
deformations of the structural system, as given in Eq. (12), where qd is the displacement
global ductility, and de,ULS is the displacement determined by a linear analysis based on the
design response spectrum, obtained from the elastic spectrum introducing the behaviour
factor q C 1.5 as a function of the design ductility class and the structural system. Dis-
placements at the damage limitation limit state ds,DLS can be estimated according to
Eq. (13), introducing the reduction coefficient m, which accounts for the lower return
period of the seismic action associated with the damage limitation requirement, in line with
general Eurocode 8 – Part 1 provisions (CEN 2004a), or as the displacement de,el,DLS
induced by a site-specific elastic spectrum defined for verifications at the damage limitation
limit state, as given in the Italian code (NTC08 2008).

123
416 Bull Earthquake Eng (2018) 16:397–425

Fig. 13 Force reduction and displacement ductility a equal displacement rule; b equal energy rule

ds;ULS ¼ qd de;ULS ð12Þ

ds;DLS ¼ mqd de;ULS  mde;el;ULS ð13Þ

Allowing a refinement of the displacement evaluation, both codes note that in general qd
is larger than q, if the fundamental period of the structure T1 is less than the corner period
of the design spectrum TC. Hence, for long-period structures, qd may be evaluated
according to the equal displacement rule (Fig. 13a) and for short-period structures
according to the equal energy rule (Fig. 13b). Such approximate estimation of the dis-
placement demand is based on the proposal of the N2 method (Fajfar 1999), established for
the nonlinear response of a bilinear single-degree-of-freedom system.

4.2 Inter-storey drifts as a function of the density-stiffness parameter

The presented results have shown that in storey j for a RC structure represented by the
stiffness parameter KS,j from Eq. (1), the inter-storey drift demand dw,j of the infilled
configuration related to the density-stiffness parameter KI,j from Eq. (7a) can be estimated
in function of the density-stiffness coefficient Cj given by Eq. (11) and the drift demand dj
of the corresponding bare structure.
As illustrated in Fig. 14a for Cj equal to 1.0 and in Fig. 14b for different values of Cj,
such correlation can be expressed using a bilinear curve, according to Eq. (14), where dm,j’
denotes the inter-storey drift capacity for the masonry infills in storey j, while the
parameter dC defines the relation of drifts between bare and infilled configurations for Cj
equal to 1.0.

Fig. 14 Relation of inter-storey drifts for bare (dj) and infilled (dw,j) structures in function of density-
stiffness coefficient Cj a definition of bilinear curve for Cj = 1.0; b variation of bilinear curve for different
values of Cj

123
Bull Earthquake Eng (2018) 16:397–425 417

8
>
< d0m;j dj
; dj  d0m;j þ dC Cj
dw;j ¼ d0m;j þ dC Cj ð14Þ
>
:d  d C ;
j C j dj [ d0m;j þ dC Cj

The relationship should provide a safe-sided approximation of expected inter-storey


drift demands for infilled configurations, ensuring that upper-bound values are obtained.
For the case studies considered within the scope of this study, a single value of inter-storey
0
drift capacity dm,j equal to 0.30%, equivalent to the allowable drift limit dDLS assigned to
the exceedance of damage limitation limit state performance requirements, has been
0
assumed. For a corresponding value of dC equal to 2/5 of dm,j (in this case, therefore, equal
to 0.12%), the bilinear curve in function of Cj has been found to describe reasonably the
average drift demands obtained based on nonlinear analyses. By definition, according to
Eq. (14), for the corner point of the bilinear curve the drift of the infilled configuration dw,j
0
always corresponds to the inter-storey drift capacity dm,j, while the related bare frame drifts
dj change depending on the parameter Cj.

4.3 Validation of proposed approach on results of nonlinear time-history


analyses

The proposed relationship given in Eq. (14) has been validated based on the results of
time-history analyses for the typical frame and wall-frame dual structures examined within
the scope of this study. Considering in the light of design verifications the overestimation
of in-plane inter-storey drift demands to be safe-sided, a selection of results is presented,
including some critical cases, for which the numerically obtained inter-storey drift
demands practically coincide with the proposed simplified prediction. Drift values corre-
sponding to storeys around mid-height of the buildings, where relatively high drift
demands are achieved, are presented. Specifically, results for two-dimensional frame
configurations, fully infilled with different infill typologies (F_T1, F_T2 and F_T3) are
shown for the 2nd storey of the 3-storey frames (Fig. 15), the 3rd storey of the 6-storey
frames (Fig. 16) and the 5th storey of the 9-storey frames (Fig. 17), for both ductility
classes at both limit states.
To illustrate the influence of the reduction of structural stiffness in elevation on the
relation of drifts for bare and infilled configurations, Fig. 18 shows numerically obtained
average drift demands for the 2nd, the 4th, the 6th and the 8th storey of 9-storey DCH
frames, fully infilled with different infill typologies (F_T1, F_T2 and F_T3) at the damage
limitation limit state.
For the theoretical case of full infill rather high values of the parameter Cj are obtained,
in particular for more flexible structural configurations and in upper storeys. In practise,

Fig. 15 Average and predicted drift demands, 3-storey 2-d frame: a j = 2, DLS; b j = 2, ULS

123
418 Bull Earthquake Eng (2018) 16:397–425

Fig. 16 Average and predicted drift demands, 6-storey 2-d frame: a j = 3, DLS; b j = 3, ULS

Fig. 17 Average and predicted drift demands, 9-storey 2-d frame: a j = 5, DLS; b j = 5, ULS

Fig. 18 Average and predicted drift demands, 9-storey 2-d frame DLS: a j = 2; b j = 4; c j = 6; d j = 8

Fig. 19 Average and predicted drift demands, 6-storey 3-d frame and wall-frame: a j = 2, DLS and ULS;
b j = 4, DLS and ULS

realistic building configurations will usually include parts without infills and infills with
perforations, resulting in lower values of Cj.
Furthermore, results obtained from analyses of three-dimensional frame and wall-frame
dual systems with different T3 infill distributions (P1_T3, P2_T3 and F_T3) are presented,
in particular, for the 2nd (Fig. 19a) and the 4th (Fig. 19b) storeys of the 6-storey DCM
wall-frame dual systems and of the DCH frame structures.
The comparison of numerical results to corresponding bilinear curves indicates that the
prediction of drift demands based on the proposed relationship in function of the density-
stiffness coefficient Cj adequately captures the variation of response. For the considered
case studies, which have shown a satisfactory infill performance, the average drift demands

123
Bull Earthquake Eng (2018) 16:397–425 419

of the infilled structures mostly result to be equal or less than the corresponding values
estimated according to Eq. (14). Hence, the proposed approach is found to be appropriate
for practical applications in the design of regular RC structures and can ensure the effective
prediction of upper bound values of expected drift demands for masonry infilled structures
depending on properties of RC structural elements, infill layouts and masonry typologies.
An exception to such observation may occur in the bottom storey of some case study
configurations, where because of a modification in the deformed shape of the structure, due
to the presence of infill, the reduction of drift demands may be less pronounced or even
negligible. Consequently, inter-storey drifts at the first storey of infilled structures may
conservatively be assumed to be equal to those of the corresponding bare configuration,
resulting in a density-stiffness coefficient Cj equal to zero.

4.4 Implications for inter-storey drift limitation in the design of RC


structures

A major application of the proposed procedure may be found in the design of new
buildings. Specifically, according to Eurocode 8-Part 1 (CEN 2004a), the damage limi-
tation requirements are satisfied if the inter-storey drift dj,DLS induced by the damage
limitation seismic action, does not exceed the inter-storey drift limit dDLS. Even though not
explicitly stated in the code, the drift verifications are commonly carried out on the bare
frame structural configuration. In fact, considering masonry infills as brittle, the require-
ment to apply the drift limit dDLS equal to 0.50% for infilled structures would imply that
rigid infills could remain substantially undamaged up to this level of drift and satisfy the
performance requirements. However, such implication is not supported by observations
based on field surveys (e.g. Braga et al. 2011) and experimental findings (e.g. Combescure
et al. 1996; Calvi and Bolognini 2001) also indicate that in traditional unreinforced infills
considerable damage is reached at significantly lower levels of drift.
The proposed prediction of inter-storey drifts for infilled structures may facilitate the
introduction of more appropriate procedures, accounting in a simplified manner for the
influence of masonry infills on the structural response. Hence, an adequate solution for the
improvement of current provisions related to infill damage control for infills built in full
contact with the surrounding RC structure can be found in the verification of inter-storey
drifts corresponding to infilled structures, rather than bare structural configurations. Inter-
storey drift verifications should be accomplished at both limit states, as illustrated in
Fig. 20.
In order to apply the proposed design approach, the elastic stiffness properties of the
load-bearing structure related to assumed dimensions of structural elements can be
expressed through the coefficient KS,j given in Eq. (1). The expected distribution of infills
within the plan of each storey, including the presence of significant openings, should be
defined based on architectural layouts. To assess the infill strength and stiffness contri-
bution to the structural response through the coefficient KI,j given in Eq. (7a), values of
masonry thickness tw,i,j and strength fw,i,j need to be known for each infill. Moreover,
characteristic values of inter-storey drift accounting for the deformation capacity of the
adopted infill typologies d0m;i;j (i.e. dDLS), which correspond to the attainment of damage
limitation limit state performance requirements, have to be defined.
Inter-storey drift limits for different limit states, related to the acceptable extent of
damage for different infill typologies, may be derived based on experimental data from in-
plane cyclic tests. Masonry infill limit states have been defined for instance in the work by

123
420 Bull Earthquake Eng (2018) 16:397–425

Fig. 20 Proposed improvements of inter-storey drift verifications in the current European design approach

Hak et al. (2013a) through the interpretation of test results (Calvi and Bolognini 2001) for
traditional unreinforced and lightly reinforced weak/slender infill typologies. Similarly, in
recent studies, extensive experimental tests have been carried out for contemporary strong
clay block masonry infills (e.g. da Porto et al. 2013; Morandi et al. 2014). They have also
resulted in the evaluation of limit states in function of inter-storey drift limits (Morandi
et al. 2014). A summary of inter-storey drifts corresponding to the attainment of opera-
tional, damage limitation and ultimate limit state performance requirements for different
infill solutions is given in Table 12. Within the scope of this study, the deformation
capacity assigned to unreinforced slender/weak infill (d0m ¼ 0:30%; du ¼ 1:00%Þ has been
conservatively assumed for all investigated infill typologies.
In function of the density-stiffness coefficient Cj, evaluated according to Eq. (11), the
bilinear curve given by Eq. (14) can be defined and expected inter-storey drift demands in
each storey of the infilled configuration can be estimated. The density-stiffness coefficient
Cj for the estimation of inter-storey drifts at the first storey should be taken equal to zero, in
order to ensure a safe-sided estimation.
Finally, inter-storey drift verifications at both limit states can be carried out. The esti-
mated drift demands of infilled structural configurations dw,j,DLS and dw,j,ULS should not
exceed the drift limits dDLS and dULS, corresponding to the attainment of damage limitation
and ultimate limit state performance requirements with reference to acceptable infill
damage. At this point, the force-based procedure that needs to be followed in order to
design the structural elements and achieve at the same time sufficient damage control for

Table 12 Inter-storey drift limits for unreinforced masonry infills at different limit states
Limit state Strong infill Slender/weak infill
(Morandi et al. 2014) (Calvi and Bolognini 2001; Hak et al. 2013a)

Operational dOLS [%] 0.30 0.20


Damage dDLS [%] 0.50 0.30
Ultimate dULS [%] 1.75 1.00

123
Bull Earthquake Eng (2018) 16:397–425 421

the displacement-sensitive non-structural masonry infills may result to be iterative, in


particular at higher levels of seismicity. An effective infill damage control should be
envisaged, ensuring that dimensions of RC structural elements are not increased with the
only purpose of limiting the damage of non-structural elements. Given that the displace-
ment demands for the infilled structure are obtained in function of building parameters, an
optimisation of the dimensions of structural elements is possible. A complete design
example can be found in Hak et al. (2013b).
In addition to the aspects considered within the scope of this study, the estimation of in-
plane drift demands for infilled RC structures may induce further developments in the
current design approach, such as improvements in the out-of-plane infill verification,
accounting for simultaneous in-plane and out-of-plane actions (Morandi et al. 2013), or in
the assessment of local effects on RC structural elements due to the presence of infills (Hak
et al. 2013c).

5 Conclusions

Motivated by the findings of previous investigations, which have demonstrated that the
current European design approach for RC structures is not always able to effectively
control the damage to traditional masonry infills, the possibility to introduce improvements
in the verification procedure based on the limitation of inter-storey drifts has been
explored. Given the fact that existing procedures provide a single drift limit for the lim-
itation of damage to masonry infills rigidly attached to the structure, applied to bare
structural configurations, independently of the structural system, infill layout and masonry
properties, the introduction of improved in-plane drift verification criteria, compliant with
specific design conditions, to be accomplished at both the damage limitation and the
ultimate limit state, is found to be essential for the effective in-plane infill damage control.
Therefore, the possibility to evaluate approximate inter-storey drift demands of infilled
RC structures in a simplified manner, based on analyses which are in principle carried out
assuming bare structural models, has been investigated. Introducing for each storey j the
density-stiffness coefficient Cj, which can been evaluated from assumed elastic structural
properties, knowing the expected infill distribution commonly available from architectural
layouts and the most relevant properties of masonry typologies to be adopted, a rela-
tionship between inter-storey drift demands of bare and infilled structural configurations
has been proposed. Validating the proposed approach on results of extensive nonlinear
numerical analyses for prototype building configurations, including frame and wall-frame
dual structural systems, different infill layouts and masonry typologies, obtained results
have demonstrated that a safe-sided prediction of inter-storey drift demands for infilled
configurations can be obtained. The density-stiffness coefficient Cj can effectively account
for the contribution of different amounts and types of infill present in the building,
reproducing adequately the reduction of drift demands for the infilled configuration in
comparison to that of the bare structure. Moreover, the influence of structural stiffness
properties on the response of the infilled configuration can be successfully captured. Based
on a comparison of deformed shapes for bare and infilled configurations, the obtained
numerical results reveal that in the ground storey of the infilled frame the reduction of drift
demands with respect to the bare configuration is less pronounced than in the upper storeys.
Hence, for the prediction of drift demands neglecting the contribution of infill in the

123
422 Bull Earthquake Eng (2018) 16:397–425

ground floor is considered to be a conservative choice. Therefore, assuming the value of Cj,
which corresponds to the ground floor, to be equal to zero is proposed.
Even though the application of the proposed procedure, considering the amount of infill
in each storey, seems to be a reasonable choice in the attempt to account for the presence of
soft-storey mechanisms, no systematic analyses have been carried out regarding the
response of buildings with irregular distributions of infill in elevation. Furthermore,
irregular distributions of infill in plan, possibly inducing a substantial contribution to the
torsional response, have not been considered within the scope of this study; hence, the
obtained results are thought to be applicable to regular building configurations without
significant irregularities regarding the distribution of infills in plan or elevation.
In conclusion, the evaluation of approximate inter-storey drift demands for infilled RC
structures, obtained without the need to perform more sophisticated analyses than in
common building design procedures, may lead to significant improvements regarding the
inter-storey drift verifications carried out to control the damage of masonry infills. In
addition, the prediction of structural drifts for infilled configurations can be applied also in
other design verifications, in particular related to combined in-plane and out-of-plane
actions and possible local effects on RC structural elements due to the presence of masonry
infill.
Even though the current force-based design approach for regular RC structures with
masonry infills, commonly adopted in design practise based on the analysis of bare frame
structural configuration, is in principle considered to be appropriate, regarding the limi-
tation of drifts, verifications accounting for the expected response considering the presence
of infill appear to be more effective. Hence, the most important advantage of the proposed
approach with respect to current seismic design recommendations can be seen in the
possibility to verify estimated drift demands of infilled structures directly against allowable
drift limits at different performance levels. Such drift limits corresponding to different
infill typologies can be evaluated experimentally by an appropriate interpretation of the
results of in-plane cyclic tests on infilled RC frames. Furthermore, the iterative nature of
the verification approach, implying a new value of the density-stiffness coefficient Cj after
each modification of the dimensions for load-bearing structural elements, limits to some
extent the increase of structural stiffness to meet the infill damage control criteria,
imposing instead a sort of optimisation regarding the relation of structural and masonry
infill properties. In order to reduce the need for iteration due to governing criteria related to
the verification of allowable displacements in the classical force-based design approach, a
possible direction of further developments lies in the evaluation of behaviour factors
q which would be more appropriate as regards the deformation capacity of typical masonry
typologies adopted for infilled RC structures. On the other hand, a successful solution may
be sought in the development of a displacement-based procedure, through the introduction
of provisions required to effectively implement this alternative design approach to infilled
RC structures.

Acknowledgements This work, conducted at the University of Pavia and at Eucentre of Pavia in Italy, was
funded by the ANDIL Assolaterizi and through the Executive Project DPC-RELUIS 2010–2013 and DPC-
RELUIS 2014–2016. The financial support received is gratefully acknowledged.

123
Bull Earthquake Eng (2018) 16:397–425 423

References
Al-Chaar G (2002) Evaluating strength and stiffness of unreinforced masonry infill structures, ERDC/CERL
TR-02-1 Research Report
Asteris PG, Antoniou ST, Sophianopoulos DS, Chrysostomou CZ (2011a) Mathematical macromodeling of
infilled frames: state of the art. J Struct Eng 137(12):1508–1517
Asteris PG, Chrysostomou CZ, Giannopoulos IP, Smyrou E (2011b) Masonry infilled reinforced concrete
frames with openings. In: Proceedings of the COMDYN, III ECCOMAS thematic conference on
computational methods in structural dynamics and earthquake engineering, Corfu, Greece
Beyer K, Bommer JJ (2007) Selection and scaling of real accelerograms for bi-directional loading: a review
of current practice and code provisions. J Earthq Eng 11(S1):13–45
Braga F, Manfredi V, Masi A, Salvatori A, Vona M (2011) Performance of non-structural elements in RC
buildings during the L’Aquila, 2009 earthquake. B Earthq Eng 9(1):307–324
Calvi GM, Bolognini D (2001) Seismic response of RC frames infilled with weakly reinforced masonry
panels. J Earthq Eng 5(2):153–185
Carr AJ (2007) Ruaumoko manual. University of Canterbury, Cristchurch
Cavaleri L, Di Trapani F (2014) Cyclic response of masonry infilled RC frames: experimental results and
simplified modelling. Soil Dyn Earthq Eng 65:224–242
CEN (2002) Eurocode 1-actions on structures, Part 1-1: general actions—densities, self-weight, imposed
loads for buildings, EN 1991-1-1. European Committee for Standardisation, Brussels
CEN (2004a) Eurocode 8-design of structures for earthquake resistance, Part 1: generalrules, seismic actions
and rules for buildings, EN 1998-1. European Committee for Standardisation, Brussels
CEN (2004b) Eurocode 2-design of concrete structures, Part 1-1: general rules and rules for buildings, EN
1992-1-1. European Committee for Standardisation, Brussels
Combescure D, Pires F, Cerqueira P, Pegon P (1996) Test on masonry infilled RC frames and its numerical
interpretation. In: Proceedings of the 11th world conference on earthquake engineering, Acapulco,
Mexico
Crisafulli FJ (1997) Seismic behaviour of reinforced concrete structures with masonry infills. Ph.D. Dis-
sertation, Department of Civil Engineering, University of Canterbury, New Zealand
da Porto F, Guidi G, Dalla Benetta M,Verlato N (2013) Combined in-plane/out-of-plane experimental
behaviour of reinforced and strengthened infill masonry walls. In: Proceedings of the 12th Canadian
Masonry Symposium, Vancouver, British Columbia
Dawe JL, Seah CK (1988) Lateral load resistance of masonry panels in flexible steel frames. In: Proceedings
of the 8th international brick and block masonry conference, Dublin, Ireland
Decanini LD, Bertoldi SH, Gavarini C (1993) Telai tamponati soggetti ad azione sismica, un modello
semplificato: confronto sperimentale e numerico. In: Atti del VI convegno nazionale ANIDIS, Perugia,
Italy (in Italian)
Decanini LD, Liberatore D, Liberatore L, Sorrentino L (2012) Preliminary Report on the 2012, May 20,
emilia earthquake, v.1. http://www.eqclearinghouse.org/2012-05-20-italy-it/
Decanini LD, Liberatore L, Mollaioli F (2014) Strength and stiffness reduction factors for infilled frames
with openings. Earthq Eng Eng Vib 13(3):437–454
Fajfar P (1999) Capacity spectrum method based on inelastic demand spectra. Earthq Eng Struct D
28(9):979–993
Fardis MN (2006) Seismic design issues for masonry-infilled RC frames. In: Proceedings of the 1st
European conference on earthquake engineering and seismology, Geneva, Switzerland
Fardis MN, Panagiotakos T (1997) Seismic design and response of bare and masonry-infilled RC buildings.
PART II: infilled Structures. J Earthq Eng 1(3):475–503
Flanagan RD, Bennett RM (1999) Bidirectional behaviour of structural clay tile infilled frames. J Struct Eng
125(3):236–244
Giberson MF (1967) The response of nonlinear multi-story structures subjected to earthquake excitation,
EERL Report. California Institute of Technology, Pasadena
Hak S, Morandi P, Magenes G, Sullivan T (2012) Damage control for clay masonry infills in the design of
RC frame structures. J Earthq Eng 16(S1):1–35
Hak S, Morandi P, Magenes G (2013a) Evaluation of infill strut properties based on in-plane cyclic tests.
Grad̄evinar 65(6):509–521
Hak S, Morandi P, Magenes G (2013b) Damage control of masonry infills in seismic Design. Report
EUCENTRE 2013/01. Iuss Press, Pavia
Hak S, Morandi P, Magenes G (2013c) Local effects in the seismic design of RC frame structures with
masonry infills. In: Proceedings of the 4th ECCOMAS thematic conference on computational methods
in structural dynamics and earthquake engineering, Kos, Greece

123
424 Bull Earthquake Eng (2018) 16:397–425

Hermanns L, Fraile A, Alarcón E, Álvarez R (2014) Performance of buildings with masonry infill walls
during the 2011 Lorca earthquake. B Earthq Eng 12(5):1977–1997
Iervolino I, Galasso C (2009) REXEL 2.31 beta-tutorial. http://www.reluis.it/doc/software/REXEL_
Tutorial_ENG.pdf
Iervolino I, Maddaloni G, Cosenza E (2008) Eurocode 8 compliant real record sets for seismic analysis of
structures. J Earthq Eng 12(1):54–90
Iervolino I, Galasso C, Cosenza E (2010) REXEL: computer aided record selection for code-based seismic
structural analysis. B Earthq Eng 8(2):339–362
Ioannou I, Borg R, Novelli V, Melo J, Alexander D, Kongar I, Verrucci E, Cahill B, Rossetto T (2012) The
29th May 2012 emilia romagna earthquake, EPICentre Field Observation Report. http://www.
eqclearinghouse.org/2012-05-20-italy/reports/
Kakaletsis DJ, Karayannis CG (2008) Influence of masonry strength and openings on infilled RC frames
under cyclic loading. J Earthq Eng 12(2):197–221
Koutromanos I, Stavridis A, Shing PB, Willam K (2011) Numerical modeling of masonry-infilled RC
frames subjected to seismic loads. Comput Struct 89(11–12):1026–1037
Luzi L, Hailemikael S, Bindi D, Pacor F, Mele F, Sabetta F (2008) ITACA (ITalian ACcelerometric
Archive): a web portal for the dissemination of italian strong-motion data. Seismol Res Lett
79(5):716–722
Magenes G, Pampanin S (2004) Seismic reponse of gravity-load design frames with masonry infills. In:
Proceedings 13th world conference on earthquake engineering, Vancouver, Canada
Magenes G, Bracchi S, Graziotti F, Mandirola M, Manzini CF, Morandi P, Palmieri M, Penna A, Rosti A,
Rota M, Tondelli M (2012) Preliminary damage survey to masonry structures after the May 2012
Emilia earthquakes, v.1. http://www.eqclearinghouse.org/2012-05-20-italy/reports/
Markulak D, Radić I, Sigmund V (2013) Cyclic testing of single bay steel frames with various types of
masonry infill. Eng Struct 51:267–277
Masi A, Vona M, Mucciarelli M (2011) Selection of natural and synthetic accelerograms for seismic
vulnerability studies on reinforced concrete frames. J Struct Eng 137(3):367–378
Mohammadi M, Akrami V, Mohammadi-Ghazi R (2011) Methods to improve infilled frame ductility.
J Struct Eng 137(6):646–653
Morandi P, Hak S, Magenes G (2011) Comportamento sismico delle tamponature in laterizio in telai in c.a.:
definizione dei livelli prestazionali e calibrazione di un modello numerico. In: Atti del XV Convegno
ANIDIS-L’ingegneria Sismica in Italia; 18–22 September 2011, Bari, Italy
Morandi P, Hak S, Magenes G (2013) Simplified out-of-plane resistance verification for slender clay
masonry infills in RC frames. In: Atti del XV Convegno ANIDIS-L’ingegneria Sismica in Italia; 30
Giugno-4 Luglio 2013, Padova, Italy
Morandi P, Hak S, Magenes G (2014) In-plane experimental response of strong masonry infills. In: Pro-
ceedings of the 9th international masonry conference, Guimarães, Portugal
Mosalam KM, Gunay S (2015) Progressive collapse analysis of RC frames with URM infill walls con-
sidering in-plane/out-of-plane interaction. Earthq Spectra 31(2):921–943
Negro P, Colombo A (1997) Irregularities induced by non-structural masonry panels in framed buildings.
Eng Struct 19(7):576–585
NTC08 (2008) Norme tecniche per le costruzioni, D.M. 14 Gennaio 2008, Ministero delle Infrastrutture,
S.O. No. 30 alla G.U. del 4.2.2008, No. 29, Rome, Italy (in Italian)
Otani S (1981) ‘‘Hysteresis models of reinforced concrete for earthquake response analysis. J Fac Eng Univ
Tokyo XXXVI(2):125–159
Pacor F, Paolucci R, Luzi L, Sabetta F, Spinelli A, Gorini A, Nicoletti M, Marcucci S, Filippi L, Dolce M
(2011) Overview of the Italian strong motion database ITACA 1.0. B Earthq Eng 9(6):1723–1739
Preti M, Bettini N, Plizzari G (2012) Infill walls with sliding joints to limit infill-frame seismic interaction:
large-scale experimental test. J Earthq Eng 16(1):125–141
Preti M, Migliorati L, Giuriani E (2015) Experimental testing of engineered masonry infill walls for post-
earthquake structural damage control. B Earthq Eng 13(7):2029–2049
Priestley MJN, Calvi GM, Kowalsky MJ (2007) Displacement-based seismic design of structures. IUSS
Press, Pavia
Ricci P, Manfredi V, De Luca F, Verderame GM (2011) 6th April 2009 L’Aquila earthquake, Italy:
reinforced concrete building performance. B Earthq Eng 9(1):285–305
Rodrigues H, Varum H, Costa A (2010) Simplified macro-model for infill masonry panels. J Earthq Eng
14(3):390–416
Shing PB, Mehrabi AB (2002) Behaviour and analysis of masonry-infilled frames. Progr Struct Eng Mater
4:320–331

123
Bull Earthquake Eng (2018) 16:397–425 425

Smyrou E, Blandon C, Antoniou S, Pinho R, Crisafulli FJ (2011) Implementation and verification of a


masonry panel model for nonlinear dynamic analysis of infilled RC frames. B Earthq Eng
9(5):1519–1534
Stavridis A, Koutromanos I, Shing PB (2012) Shake-table tests of a three-story reinforced concrete frame
with masonry infill walls. Earthq Eng Struct D 41(6):1089–1108
Sullivan TJ, Priestley MJN, Calvi GM (2006) Seismic design of frame-wall structures, Research Report
ROSE-2006/02. IUSS Press, Pavia
Vicente R, Rodrigues H, Costa A, Varum H, Mendes da Silva JAR (2012) Performance of masonry
enclosure walls: lessons learned from recent earthquakes. Earthq Eng Eng Vib 11(1):23–34

123

Vous aimerez peut-être aussi