You are on page 1of 87

 

Seismic  Vulnerability  Assessment  

 
 
 
 

Vitor  Silva  
 
 
 
 
 
May  2015  
Seismic  Vulnerability  Assessment  Course  
 
Lecturer:  Vitor  Silva  
 
Description:  
 
Earthquake  loss  estimation  can  play  a  fundamental  role  in  the  sustainable  development  of  a  
given   region,   providing   local   governments   and   other   decision   makers   with   valuable  
information   necessary   for   the   creation   of   risk   mitigation   actions.   An   important   component  
for   this   purpose   is   a   vulnerability   model   that   allows   the   estimation   of   losses   from  
structural/non-­‐structural  damage  due  to  earthquakes,  as  a  function  of  a  set  of  ground  motion  
parameters.   Structural   vulnerability   can   thus   be   defined   as   the   likelihood   of   a   certain  
element  to  suffer  loss  due  to  the  effects  of  an  earthquake.  The  influence  of  the  vulnerability  
of   the   exposed   elements   to   seismic   events   is   fundamental   in   the   magnitude   of   losses.   A  
simple  comparison  between  earthquakes  that  occurred  in  developing  regions  and  developed  
countries  reveals  the  critical  importance  of  vulnerability.  For  example,  the  Spitak  (Armenia)  
earthquake  of  December  1988  had  a  magnitude  of  Ms  6.7  and  left  a  death  toll  of  about  25000  
casualties.   Less   than   a   year   after,   an   earthquake   with   a   greater   magnitude   (Ms   7.0)   occurred  
in   Loma   Prieta   (California,   USA)   causing   a   number   of   human   losses   smaller   than   70.   This  
way,   the   structural   vulnerability   assumes   special   importance   in   the   estimation   of   seismic  
risk,  not  only  for  reflecting  directly  the  damage  susceptibility  of  a  structure,  but  also  because  
by  intervening  with  appropriate  strengthening  solutions,  it  may  be  possible  to  significantly  
reduce  the  vulnerability,  and  consequently  the  potential  for  human  losses.  
 
This   course   comprises   a   brief   introduction   to   seismic   hazard,   risk   and   exposure   modelling,  
within  the  scope  of  vulnerability  analyses.  These  notes  leverage  upon  existing  studies  from  a  
number   of   sources   that   are   cited   and   acknowledged   throughout,   and   its   structure   and  
contents   are   strongly   based   on   the   Seismic   Risk   and   Vulnerability   course   from   the  
Understanding  and  Managing  Extremes  School  at  Pavia,  Italy.  
 
Course  Schedule:  

7th  of  May  


-­‐ Basic  concepts  of  seismic  hazard,  risk  and  exposure  modelling.  
-­‐ The  importance  of  seismic  vulnerability  assessment  
-­‐ Overview  of  empirical,  analytical  and  expert  opinion  methodologies  
-­‐ Questionnaire    
 
8th  of  May  
-­‐ EMS-­‐98  approach  (Lagomarsino and Giovinazzi  2006)    
-­‐ Nonlinear  static  procedures  in  fragility  assessment  
o SPO2IDA  (Vamvatsikos  and  Cornell  2006)  
o Capacity  Spectrum  Method  (HAZUS  -­‐  FEMA  2004)  
o Displacement-­‐based  methodologies  (Silva  et  al.  2013)  
-­‐ Nonlinear  time  history  analysis  (MDOF  and  SDOF)  
-­‐ Derivation  of  vulnerability  functions  
-­‐ Exercises  
 
11th/12th  of  May  
-­‐ Presentation  of  the  Global  Earthquake  Model  
-­‐ Seismic  Vulnerability  Assessment  in  South  America  (SARA  Project)  
-­‐ Examples  of  seismic  vulnerability  and  risk  assessment  
-­‐ Questionnaire    
CHAPTER  1  INTRODUCTION  .................................................................................................  1  
1.1.   THE  NEED  FOR  EARTHQUAKE  LOSS  MODELLING  ............................................................  1  
1.2.   THE  IMPORTANCE  OF  SEISMIC  VULNERABILITY  ..............................................................  4  

CHAPTER  2  BASIC  CONCEPTS  OF  SEISMIC  HAZARD  .....................................................  6  


2.1.   GROUND  SHAKING  INTENSITY  MEASURES  .......................................................................  8  
2.1.1.   OBSERVATION-­‐BASED  INTENSITY  ...............................................................................................  9  
2.1.2.   INSTRUMENTAL  INTENSITY  ........................................................................................................  11  
2.1.2.1.   Peak  Ground  Acceleration  ...............................................................................................  11  
2.1.2.1.   Arias  Intensity  .......................................................................................................................  11  
2.1.2.2.   Significant  Durations  ..........................................................................................................  12  
2.1.2.3.   Pseudo-­‐spectral  acceleration  .........................................................................................  12  
2.2.   GROUND  MOTION  AND  INTENSITY  PREDICTION  EQUATIONS  ......................................  13  
2.3.   SHAKEMAPS  ....................................................................................................................  15  
2.4.   GROUND  MOTION  FIELDS  ..............................................................................................  17  
2.5.   PROBABILISTIC  SEISMIC  HAZARD  ASSESSMENT  ...........................................................  18  

CHAPTER  3  BASIC  CONCEPTS  OF  EXPOSURE  MODELLING  ......................................  21  


3.1.   GLOBAL  EXPOSURE  MODELLING  .....................................................................................  21  
3.2.   REGIONAL  EXPOSURE  MODELLING  .................................................................................  24  
3.3.   THE  IMPORTANCE  OF  BUILDING  TAXONOMIES  .............................................................  25  
3.4.   GEM  BASIC  BUILDING  TAXONOMY  ...............................................................................  26  

CHAPTER  4  SEISMIC  VULNERABILITY  ASSESSMENT  .................................................  28  


4.1.   MAIN  OUTPUTS  IN  VULNERABILITY  ASSESSMENT  .........................................................  28  
4.2.   EXPERT  OPINION  VULNERABILITY  ................................................................................  30  
4.3.   EMPIRICAL  VULNERABILITY  ..........................................................................................  32  
4.4.   FRAGILITY  ASSESSMENT  ................................................................................................  34  
4.4.1.   DAMAGE  STATES  ..........................................................................................................................  34  
4.4.2.   FRAGILITY  FUNCTIONS  AND  DAMAGE  PROBABILITY  MATRICES  .........................................  35  
4.4.3.   EXPERT  OPINION  FRAGILITY  .....................................................................................................  36  
4.4.4.   EMPIRICAL  FRAGILITY  .................................................................................................................  41  
4.4.5.   ANALYTICAL  FRAGILITY  ..............................................................................................................  44  
4.4.6.   NONLINEAR  STATIC  ANALYSIS  ...................................................................................................  46  
4.4.6.1.   SPO2IDA  (Vamvatsikos  and  Cornel,  2006)  ...............................................................  50  
4.4.6.2.   The  Capacity  Spectrum  Method  (HAZUS  -­‐  FEMA,  2005)  ....................................  52  
4.4.6.1.   DBELA  (Silva  et  al.  2013)  .................................................................................................  56  
4.4.7.   NONLINEAR  DYNAMIC  ANALYSIS  ...............................................................................................  62  
4.5.   CONSEQUENCE  FUNCTIONS  ............................................................................................  64  

CHAPTER  5  BASIC  CONCEPTS  OF  SEISMIC  RISK  ..........................................................  68  


5.1.   SCENARIO  RISK  ASSESSMENT  ........................................................................................  68  
5.2.   SCENARIO  DAMAGE  ASSESSMENT  ..................................................................................  69  
5.3.   PSHA  AND  EVENT-­‐BASED  RISK  ASSESSMENT  ..............................................................  70  
5.4.   RETROFITTING  DECISION  SUPPORT  TOOL  ...................................................................  72  
 
 
Chapter  1  
Introduction  
1.1. The  Need  for  Earthquake  Loss  Modelling  
In   2011   the   world   celebrated   the   birth   of   the   7   billionth   citizen.   It   has   been  
estimated   that   in   10   years   the   world   population   will   reach   8.1   billions,   and   9.4  
billions   in   the   year   of   2050   (PRB,   2010).   This   uncontrolled   growth   of   the   population  
has   led   to   an   increase   of   megacities   (with   a   population   greater   than   2   million),   often  
located  in  areas  prone  to  natural  disasters,  such  as  earthquakes.  This  peril  has  been  
responsible  for  a  death  toll  of  over  60  thousand  people  per  year  in  the  last  decades  
and  economic  losses  that  can  reach  a  great  fraction  of  a  country’s  welfare.  In  the  last  
50   years   in   Central   America,   the   earthquakes   of   Guatemala   (1976),   Nicaragua  
(1972),   and   El   Salvador   (1986)   caused   economic   losses   of   approximately   98%,   82%  
and   40%   of   the   nominal   gross   domestic   product   (GDP)   of   each   country,   respectively  
(Daniell   et   al.,   2010).   In   the   Haiti   earthquake   of   2010,   the   economic   losses   were  
above   the   nominal   GDP   (120%)   and   more   than   300   thousand   people   are   believed   to  
have   perished.   In   addition   to   these   direct   consequences   in   the   vicinity   of   the   seismic  
event,   business   disruption   of   multi-­‐national   enterprises   can   induce   a   negative  
impact  at  a  global  scale.    
The  global  economic  losses  and  insured  losses  due  to  great  natural  disasters  from  
1980  to  2011  are  illustrated  in  Figure  1.1.  
Introduction  
 

 
Figure  1.1.  Overall  losses  and  insured  losses  from  1980  until  2011  (MunichRe,  2012).  

Earthquakes  constitute  on  average  20%  of  the  overall  losses,  but  in  some  years,  
this  portion  can  be  as  high  as  60%  (e.g.  2010,  2011).  Despite  the  great  advances  that  
have   been   made   in   the   last   decades   in   the   areas   of   probabilistic   seismic   hazard  
assessment  (e.g.  Abrahamson  2006;  Bommer  and  Abrahamson  2006),  evaluation  of  
building   seismic   vulnerability   (e.g.   Calvi   et   al.   2006)   and   collection   of   information  
regarding   the   elements   exposed   to   the   hazards   (e.g.   Gamba   et   al.,   2012),   an   increase  
in   the   trend   of   earthquake   losses   is   still   observed.   The   Global   Assessment   Report  
(UNISDR,   2009)   points   out   two   probable   causes   for   this   tendency.   Firstly,   the  
significant  increase  of  population  and  capital  stock  in  hazard  prone  areas  that  have  a  
direct  impact  in  the  associated  catastrophe  risk.  Figure  1.2  illustrates  the  amount  of  
population  and  GDP  exposed  to  hazard  in  the  20  countries  with  the  highest  values.  

 
Figure  1.2.  Absolute  population  and  GDP  exposed  to  natural  disasters  (adapted  from  UNISDR,  2009).  

2
Chapter  1  

Secondly,   besides   this   growth   in   the   property   and   property   value,   modern  
societies   strongly   rely   on   inter-­‐related   systems   (building   stock,   power,   finance,  
transport),   which   in   case   of   a   natural   catastrophe   might   initiate   a   cascade   effect,  
where  a  disaster  triggers  another  disaster.  Modelling  this  system  of  systems  might  
be  very  challenging  due  to  the  required  understanding  of  each  component,  and  lack  
of  numerical  tools  to  carry  out  such  calculations.  
The   Great   East   Japan   earthquake   (2011)   and   tsunami   sent   a   clear   message   that  
both   developed   and   developing   countries   are   exposed   to   high   risks.   The   year   of  
2011  was  the  most  expensive  year  ever  registered,  far  exceeding  the  2005  economic  
losses   (hurricane   Katrina),   which   yield   the   previous   record.   From   the   overall   cost   of  
380  billion  US$,  the  earthquake  disasters  in  Japan  and  New  Zeeland  alone  accounted  
for  absolute  losses  of  228  billion  US$.    Such  losses  can  have  a  crippling  effect  in  the  
economy   of   countries   whose   governments   have   the   legal   liability   to   cover   the   full  
costs  of  rebuilding.  In  the  Kocaeli  and  Düzce  earthquakes  of  1999  (with  a  combined  
fraction  of  GDP  loss  of  approximately  8%),  the  Turkish  government  was  faced  with  
an  enormous  financial  burden  due  to  its  statuary  obligation  in  covering  the  costs  of  
reconstruction.   This   situation   propelled   the   creation   of   the   Turkish   Catastrophe  
Insurance   Pool   (TCIP),   which   allowed   transferring   large   parts   of   the   financial  
burden   due   to   seismic   losses   to   the   world’s   reinsurance   market   (Bommer   et   al.  
2002).  In  this  project,  the  creation  of  an  earthquake  loss  model  was  fundamental  to  
the  development  of  the  economic  model  used  to  evaluate  the  impact  of  catastrophe  
risk   in   the   Turkish   economy.   Furthermore,   earthquake   loss   modelling   also   serves   as  
the   foundation   to   many   other   seismic   risk   mitigation   actions.   These   may   include  
prioritization  of  zones  within  a  country  where  the  structural  seismic  vulnerability  of  
the   building   stock   should   be   improved,   planning   of   post-­‐disaster   emergency  
response  or  definition  of  regulations  to  impose  seismic-­‐proof  construction  practices.  
However,   in   less   developed   countries,   the   required   resources,   datasets   and   tools  
might   not   exist   in   order   to   perform   a   comprehensive   assessment   of   the   seismic   risk.  
In  fact,  the  evaluation  of  the  annual  costs  of  natural  disasters  from  the  last  decades  
between   low/middle-­‐income   countries   and   high-­‐income   countries   shows  
considerably  lower  costs  in  the  latter  category,  where  catastrophe  models  are  more  
frequently  available.  This  trend  can  be  seen  in  Figure  1.3.  

3
Introduction  
 

 
Figure  1.3.  Distribution  of  average  annual  cost  of  natural  disaster  in  low-­‐,  middle-­‐  and  high-­‐income  
countries  (adapted  from  Cummins  and  Mahul,  2009).  

1.2. The  importance  of  Seismic  Vulnerability    


Structural   vulnerability   can   be   defined   as   the   likelihood   of   a   certain   element   to  
suffer  loss  due  to  the  effects  of  an  earthquake.  The  influence  of  the  vulnerability  of  
the   exposed   elements   to   seismic   events   is   fundamental   in   the   magnitude   of   losses.   A  
simple   comparison   between   earthquakes   that   occurred   in   developing   regions   and  
developed   countries   reveals   the   critical   importance   of   vulnerability.   For   example,  
the  Spitak  (Armenia)  earthquake  of  December  1988  had  a  magnitude  of  Ms  6.7  and  
left  a  death  toll  of  about  25000  casualties.  Less  than  a  year  after,  an  earthquake  with  
a   greater   magnitude   (Ms   7.0)   occurred   in   Loma   Prieta   (California,   USA)   causing   a  
number   of   human   losses   smaller   than   70   (Bommer,   2004).   For   this   reason,   the  
structural   vulnerability   assumes   special   importance   in   the   estimation   of   seismic  
risk,   not   only   for   reflecting   directly   the   damage   susceptibility   of   a   structure,   but   also  
because   by   intervening   with   appropriate   strengthening   solutions,   it   may   be   possible  
to  significantly  reduce  the  vulnerability,  and  consequently  the  seismic  risk.    
The   recognition   of   the   importance   in   understanding   the   vulnerability   of   the  
exposed   elements   led   to   a   rapid   rise   in   demand   for   accurate   and   flexible  
methodologies   for   its   evaluation   (Calvi   et   al.   2006).   These   may   include   empirical  
methods   that   take   advantage   of   post-­‐earthquake   damage   data   to   derive   fragility  
functions   (e.g.   Colombi   et   al.,   2008,   Rota   et   al.   2008);   simplified   approaches   where   a  
set  of  structural  parameters  are  employed  to  derive  a  vulnerability  index,  which  is  
then   used   to   calculate   a   curve   relating   levels   of   damage   with   a   set   of   intensity  
measure   levels   (Lagomarsino   and   Giovinazzi,   2006);   or   analytical   methodologies  

4
Chapter  1  

that   rely   on   numerical   models   to   simulate   the   seismic   performance   of   structures  


against   increasing   levels   of   ground   motion   (e.g.   Vamvatsikos   and   Cornell   2002;  
Freeman   2004;   Rossetto   and   Elnashai   2005).   The   latter   category   of   methods   has   the  
advantage   of   not   depending   on   the   availability   of   post-­‐earthquake   damage   data,   and  
depending  on  the  level  of  complexity  of  the  numerical  models,  these  approaches  can  
still   incorporate   fundamental   structural   characteristics   in   the   analysis   such   as  
vertical  or  plan  irregularities  or  influence  of  higher  modes  of  vibration  (Chopra  and  
Goel   2002;   Casarotti   and   Pinho   2007).   Each   analytical   method   considers   different  
simplifications,   assumptions   and   algorithms,   mainly   in   the   way   the   structural  
nonlinearity   is   handled.   Consequently,   the   structural   response   will   be   dependent   on  
the   chosen   methodology,   and   discrepancies   in   the   order   of   2   can   be   observed  
(Chopra  and  Goel,  2000;  Lin  et  al,  2004).  For  these  reasons,  it  is  fundamental  to  fully  
comprehend   the   limitations   and   strengths   of   each   vulnerability   methodology,   and  
ensure   that   the   selected   approach   is   capable   of   endorsing   the   accuracy   and  
reliability  required  by  the  end  users.  

 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 

5
Chapter  2  
Basic  concepts  of  seismic  hazard  
Seismic   hazard   can   be   described   as   the   physical   phenomenon   induced   by   an  
earthquake   that   may   cause   loss   of   life   or   injury,   property   damage,   social   and  
economic   disruption   or   environmental   degradation.   The   physical   phenomena  
caused  by  earthquakes  are  summarized  in  Figure  2.1,  and  include  tsunamis,  surface  
rupture,  landslides,  liquefaction  and  amplified  ground  shaking.    

SURFACE FAULTING TSUNAMIS


RUPTURE

SEISMIC
ENERGY
RELEASE

LANDSLIDES LIQUEFACTION

TOPOGRAPHIC STRONG
SOIL DEPOSITS
RELIEF GROUND
MOTION

AMPLIFIED AMPLIFIED
GROUND GROUND
SHAKING SHAKING
 
Figure  2.1.  The  various  natural  phenomenon  caused  by  the  release  of  seismic  energy.  

The   various   phenomena   caused   by   earthquakes   will   produce   damage   in   different  


manners   in   the   surrounding   building   inventory,   and   it   is   fundamental   to   fully  
comprehend   their   physical   behavior   in   order   to   properly   assess   the   seismic  
Chapter  3  

vulnerability  of  the  built  environment.  The  following  figures  demonstrate  the  effects  
of  these  different  phenomena.    

   
Figure  2.2.  Fault  rupture.  

     
Figure  2.3.  Tsunamis.  

   
Figure  2.4.  Landslides.  

   
Figure  2.5.  Liquefaction.  

7
Basic concepts of seismic hazard

The  different  natural  phenomenon  induced  in  earthquakes  from  1989  to  2003  and  
the  associated  damage  was  studied  by  Bird  and  Bommer  (2004).  The  main  objective  
of   the   later   study   was   to   identify   the   main   and   secondary   causes   of   damage   to  
buildings,   transport   systems   (e.g.   roads   and   railway   lines)   and   utilities   (e.g.  
pipelines).  Figure  2.6  shows  the  results  considering  only  building  damage.  

Tsunami Shaking
Landslide
Landslide

None
Liquefaction

Tsunami
Shaking Fault rupture

Figure  2.6.  The  primary  cause  (left)  and  secondary  cause  of  damage  to  buildings  from  50  destructive  
earthquakes  from  1989-­‐2003  (Bird  and  Bommer,  2004).  

Bird   and   Bommer   (2004)   concluded   that   there   were   very   few   cases   where   ground  
shaking  did  not  dominate  on  a  regional  scale,  but  in  smaller  survey  areas  secondary  
hazards   were   seen   to   also   contribute   significantly   (e.g.   in   the   Turkish   town   of  
Adapazari,   extensive   liquefaction   was   observed   durig   the   1999   Kocaeli   earthquake).  
The  majority  of  the  existing  vulnerability  models  thus  focus  on  ground  shaking.  

2.1. Ground  Shaking  Intensity  Measures  


An   earthquake   can   be   defined   as   a   sudden   release   of   elastic   strain   energy   in   the  
Earth’s  crust  as  the  result  of  a  fault  rupture  (source).  Energy  radiates  from  the  fault  
rupture   and   from   the   surrounding   volume   of   the   crust   in   the   form   of   elastic   or  
seismic   waves   (path).   Strong   ground   shaking   occurs   when   the   seismic   waves   pass  
through  soil  layers  (which  cause  refraction  and  amplification  of  the  seismic  waves)  
and   reach   the   surface   of   the   earth   (site).   The   intensity   of   ground   shaking   from   an  
earthquake   at   a   given   site   can   be   measured   using   a   number   of   parameters.   These  
can  be  categorized  into  two  types:  intensity  that  is  assigned  based  on   observations  
(also   known   as   macroseismic   intensity);   and   intensity   that   is   measured   from  
instruments,  such  as  accelerographs.    

8
Chapter  3  

2.1.1. Observation-­‐based  Intensity  


Bolt   (1999)   defines   intensity   as   “a   measure   of   ground   shaking   obtained   from  
damage  done  to  structures  built  by  humans,  changes  in  the  Earth’s  surface,  and  felt  
reports”.   Intensity   is   an   index   that   reflects   the   strength   of   ground   shaking   at   a  
particular   location   during   an   earthquake.   Therefore,   it   is   not   really   a   measure   of   the  
size   of   the   earthquake   in   the   same   way   as   moment   or   magnitude,   but   rather   a  
measure   of   ground   motion.   In   order   to   make   clear   the   index   nature   of   intensity   it  
was   historically   represented   by   Roman   numerals,   but   modern   practice   has   moved  
towards   the   use   of   Arabic   numerals   to   ease   their   use   in   risk   algorithms   and  
numerical  methodologies.    

Intensity   scales   were   initially   proposed   by   Rossi   and   Forel   in   1883.   The   historical  
evolution   of   intensity   scales   is   described   by   Coburn   and   Spence   (2002).   Many   of   the  
earlier  scales  have  now  been  completely  abandoned  and  the  main  scales  are  those  of  
the  Modified  Mercalli  (MM)  scale,  used  in  the  Americas,  and  the  Medvedev-­‐Karnik-­‐
Sponheuer   (MSK)   scale   used   in   most   of   Europe,   except   Italy   where   the   Mercalli-­‐
Cancani-­‐Sieberg   scale   (MCS)   is   employed.   The   MSK   scale   has   now   generally   been  
replaced  by  the  European  Macroseismic  Scale  (Grünthal,  1998)  (EMS-­‐98).  The  MM,  
MSK   and   EMS   scales,   as   well   as   others,   base   the   grades   of   intensity   on   (a)   the  
shaking   felt   by   humans,   (b)   the   movement   of   objects,   (c)   the   damage   to   buildings,  
and  (d)  changes  in  the  ground.  This  last  category  is  not  a  reliable  indicator  because  
of  the  very  high  heterogeneity  of  soils  and  rocks  and  the  strong  influence  of  ground  
water   conditions   and   other   parameters.   It   is   generally   advised   not   to   use  
observations  of  ground  behavior  other  than  for  corroboration  with  the  MM  and  MSK  
scales.   In   fact,   the   EMS-­‐98   removed   the   geotechnical   and   geological   observations  
from  the  descriptions  of  different  grades  of  intensity.    

The  lower  levels  of  intensity  are  defined  primarily  by  how  people  feel  the  shaking,  
but  as  seismic  intensity  becomes  more  severe,  human  perception  of  the  movement  
becomes  progressively  less  important,  and  damage  to  buildings  assumes  a  dominant  
role.   Intensity   III   is   generally   considered   as   the   threshold   of   perceptibility,   below  
which  the  ground  shaking  is  not  felt  by  most  people.  Intensity  VII  can  be  thought  of  
as   the   threshold   of   appreciable   building   damage,   although   this   might   be   intensity  
VIII   for   engineered   structures.   Intensity   XII   is   very   rarely,   if   ever,   encountered   in  

9
Basic concepts of seismic hazard

reality  and  therefore  XI  can  be  treated  as  the  upper  bound.  In  practice,  X  appears  to  
be  an  effective  upper  bound.  

One   final   point   must   be   made:   the   scales   are   neither   continuous   nor   linear.   The  
variation   from   degree   to   degree   is   not   gradual,   each   increase   in   one   degree  
representing   a   jump   in   the   level   of   shaking.   Furthermore,   the   jumps   between  
different  degrees  are  not  equal:  the  increase  in  the  level  of  ground  shaking  from  IV  
to  V  is  not  the  same  as  the  increase  from  VII  to  VIII.    

One  of  the  most  successful  and  well-­‐established  efforts  to  calculate  and  disseminate  
observational   ground   shaking   intensity   shortly   after   the   occurrence   of   an  
earthquake  is  the  Did  You  Feel  it?  initiative.  This  service  has  been  supported  by  the  
United   States   Geological   Survey   (USGS)   to   produce   Community   Internet   Intensity  
Maps   (CCIM),   which   summarize   the   questionnaire   responses   provided   by   internet  
users.    “Communities”  are  defined  as  zip  code,  or  postal  code,  regions.  The  form  of  
the   questionnaire   and   the   method   for   assignment   of   intensities   are   based   on   an  
algorithm   developed   by   Dengler   and   Dewey   (1998)   for   determining   a   ”Community  
Decimal  Intensity".  Figure  2.7  shows  the  CIIM  for  an  earthquake  in  Northern  Italy.    

 
Figure  2.7.  Community  Internet  Intensity  Map  for  the  M5.2  earthquake  in  Northern  Italy  on  27th  
January  2012  

10
Chapter  3  

2.1.2. Instrumental  Intensity  


The   instruments   that   record   the   acceleration   of   the   ground   as   a   function   of   time  
strong   ground   motion   are   called   accelerographs   (unlike   seismographs   which   record  
the   displacement   or   velocity   of   the   ground).   The   first   accelerographs   were  
developed  and  installed  in  California  in  1932,  more  than  three  decades  after  the  first  
seismographs   came   into   operation.   Accelerograms,   the   records   obtained   from  
accelerographs,   contain   a   wealth   of   information   about   the   nature   of   the   ground  
shaking  in  strong  earthquakes  and  also  about  the  highly  varied  characteristics  that  
different  earthquakes  can  produce  at  different  locations.  

2.1.2.1. Peak  Ground  Acceleration  


As  shown  in  Figure  2.8,  peak  ground  acceleration  (PGA)  relates  only  to  one  isolated  
peak  within  a  record  of  ground  motion.  Therefore,  it  is  a  rather  poor  parameter  for  
characterizing   the   entire   ground   motion   record.   The   numerical   integration   of   the  
accelerogram  allows  the  velocity  and  displacement  time-­‐histories  to  be  calculated  as  
well,  from  which  peak  ground  velocity  (PGV)  and  peak  ground  displacement  (PGD)  
can  be  obtained.    

Figure  2.8.  Representation  of  the  peak  ground  acceleration  (PGA)  in  a  ground  motion  record.  

2.1.2.1. Arias  Intensity  


Arias   Intensity   (AI)   is   another   measure   of   the   energy   of   the   ground   motion   record  
and  it  is  defined  as  the  time-­‐integral  of  the  square  of  the  ground  acceleration:  

!!
𝜋
𝐼! = 𝑎(𝑡)! 𝑑𝑡  
2𝑔 !

11
Basic concepts of seismic hazard

where   g   is   acceleration   due   to   gravity   and   Td   is   the   duration   of   the   accelerogram  


above  a  given  acceleration  threshold.  

Figure  2.9.  Representation  of  cumulative  Arias  Intensity  versus  time.  

2.1.2.2. Significant  Durations  


The   significant   duration   is   the   time   over   which   significant   ground   shaking   occurs.  
The   Arias   Intensity   can   be   cumulatively   calculated   with   time   (as   shown   in   the   figure  
above),   and   significant   durations   can   be   defined   which   describe   the   length   of   time  
over  which  the  Cumulative  Arias  Intensity  is  between  5%  and  75%  of  IA,  or  between  
5%  and  95%  of  IA.    

2.1.2.3. Pseudo-­‐spectral  acceleration  


If   a   number   of   single   degree   of   freedom   (SDOF)   systems   (i.e.   equivalent   to   a   number  
masses   fixed   on   top   of   several   sticks   with   distinct   heights   –   see   Figure   2.10)   are  
subjected   to   a   given   acceleration   time-­‐history   at   their   base,   they   will   respond  
differently.   If   the   peak   response   of   each   SDOF   system   (acceleration,   velocity,  
displacement)   is   registered,   and   plotted   against   its   natural   period   of   vibration,   a  
response  spectrum  is  produced.    

Figure  2.10.  Representation  of  a  number  of  single  degree  of  freedom  (SDOF)  systems.  

12
Chapter  3  

 
Figure  2.11.  Acceleration  response  spectra  for  three  elastic  damping  levels.  

2.2. Ground  Motion  and  Intensity  Prediction  


Equations  
Given  a  large  number  of  records,  one  can  calculate  values  for  any  of  the  parameters  
described  in  the  previous  section  and  obtain  a  robust  estimate  of  the  correlation  of  
these  values  with  any  other  parameter  relevant  to  this  suite  of  records,  such  as  the  
magnitude   of   the   earthquake   from   which   they   came.   This   type   of   reasoning   is   the  
basis   for   the   development   of   empirical   predictive   equations   for   strong   ground  
motions.   Usually,   a   relationship   is   sought   between   a   suite   of   observed   ground  
motion   parameters   and   an   associated   set   of   independent   variables   including   a  
measure  of  the  size  of  the  earthquake,  a  measure  of  distance  from  the  source  to  the  
site,   some   classification   of   the   style-­‐of-­‐faulting   involved   and   some   description   of   the  
geological   and   geotechnical   conditions   at   the   recording   site.   An   empirical   ground  
motion   prediction   equation   (GMPE)   is   simply   a   function   of   these   independent  
variables   that   provides   an   estimate   of   the   expected   value   of   the   ground   motion  
parameter   in   consideration,   as   well   as   some   measure   of   the   distribution   of   values  
about  this  expected  value.  

Thus  far  the  development  of  empirical  ground  motion  prediction  equations  has  been  
almost  exclusively  focused  upon  the  prediction  of  peak  ground-­‐motions,  particularly  
PGA  and,  to  a  far  lesser  extent,  PGV,  and  ordinates  of  5%  damped  elastic  acceleration  
response   spectra   (Douglas,   2003).   Regardless   of   the   ground   motion   measure   in  
consideration,   a   GMPE   can   be   represented   as   a   generic   function   of   predictor  
variables,  μ(M,R,θ)  and  a  variance  term,  εσT,  as  in  the  following  equation:  

log 𝑦 = 𝜇 𝑀, 𝑅, 𝜃 + 𝜀𝜎!  

13
Basic concepts of seismic hazard

For   most   ground   motion   measures,   the   values   will   increase   with   increasing  
magnitude  and  decrease  with  increasing  distance.  These  two  scaling  effects  form  the  
backbone  of  prediction  equations  and  many  functional  forms  have  been  proposed  to  
capture   the   variation   of   motions   with   respect   to   these   two   predictors   (Douglas,  
2003).   Different   magnitude   measures   and   different   distance   terms,   such   as  
hypocentral   distance   and   Joyner-­‐Boore   distance   can   be   used   in   GMPEs,   though   the  
most   commonly   used   in   recent   studies   are   Moment   Magnitude   (MW)   and   Joyner-­‐
Boore   distance   (rJB).   For   any   particular   ground   motion   record,   the   total   variance  
term  given  in  the  equation  above  may  be  partitioned  into  two  components  as  shown  
below:    

log 𝑦!" = 𝜇 𝑚!" , 𝑟!" , 𝜃!" + 𝛿!" + 𝛿!,!!  

The   terms   δe,i   and   δa,ij   represent   the   inter-­‐event   and   intra-­‐event   residuals,  
respectively,  and  quantify  how  far  away  from  the  mean  estimate  of  logyij  the  motions  
from   the   ith   event   and   the   jth   recording   from   the   ith   event   are,   respectively,   as  
illustrated  in  Figure  2.12.    

 
Figure  2.12.  The  predicted  median  ground  motion  parameter  varying  with  distance  (bold  line),  the  
actual  recordings  on  which  they  are  based  for  two  different  events  (circle  and  square),  and  the  
representation  of  inter-­‐  and  intra-­‐event  residuals  (Bommer  and  Stafford  2008)  

Equations  to  estimate  the  macroseismic  intensity  for  a  region  given  magnitude  and  
distance  are  also  available,  often  being  derived  in  regions  where  macroseismic  data  
is   much   more   abundant   than   instrumental   intensity   (due   to   a   lack   of   recording  

14
Chapter  3  

stations).  These  so-­‐called  Intensity  Prediction  Equations  (IPEs)  often  have  the  same  
format  as  that  presented  previously  for  instrumental  ground  motion  parameters.  A  
review  of  various  IPEs  is  available  in  Cua  et  al.  (2010).    

 
Figure  2.13.  Variation  in  intensity  predictions  from  various  IPEs  (Cua  et  al.  2010)  

2.3. ShakeMaps  
As   defined   by   Wald   et   al.   (1999),   a   shake   map   is   a   representation   of   ground   shaking  
produced   by   an   earthquake.   In   order   to   generate   rapid-­‐response   ground   motion  
maps   it   is   necessary   to   determine   the   best   format   for   reliable   presentation   of   the  
maps  given  the  diverse  audience,  such  as  scientists,  businesses,  emergency  response  
agencies,  media,  and  the  general  public.  As  discussed  in  Wald  et  al.  (1999),  the  use  of  
intensity  maps  instead  of  peak  ground  acceleration  and  velocity  maps  could  help  to  
simplify  and  maximize  the  flow  of  information  to  the  public.  

ShakeMap   is   a   product   of   the   United   States   Geological   Survey   (USGS)   Earthquake  


Hazards   Program   in   union   with   regional   seismic   network   operators.     This   project  
makes   available   near-­‐real-­‐time   maps   of   ground   motion   and   shaking   intensity  
following   significant   earthquakes.   The   methodology   applied   for   generating   these  
maps  can  be  summarized  as  follows:    

1. Gather  the  observed  peak  ground  motions  at  recording  stations  in  the  region  
of  the  earthquake  (these  are  shown  as  triangles  on  the  maps).  

15
Basic concepts of seismic hazard

2. Determine   the   centroid   (magnitude   and   location)   of   the   event   and   assess  
median   ground   motions   on   rock   at   sites   far   from   recording   stations   (using  
GMPEs).  

3. Correct   the   recorded   data   to   rock   based   on   the   site   QTM   (Quaternary   Tertiary  
Mesozoic)  geology  and  interpolate  the  data  plus  the  predicted  ground  motions  
onto  a  fine  rock  grid.  

4. Site   amplify   at   each   fine   grid   point   based   on   its   site   QTM   geology   and  
amplitude,  fit  a  smooth  function  through  all  fine  grid  points,  and  contour.  

It   is   important   to   note   that   most   of   the   areas   that   could   be   affected   by   a   large  
earthquake  do  not  have  sufficient  instrumentation  for  recording  the  ground  motion  
parameters   associated   with   the   seismic   event.     Hence,   there   are   significant   gaps   in  
the   observed   shaking   distribution   which   makes   it   necessary   to   develop   algorithms  
to   best   describe   the   shaking   in   more   remote   areas   by   utilizing   a   variety   of  
seismological  tools.  

As  described  by  Wald  et  al.  (1999),  to  create  the  best  composite  map  it  is  necessary  
to   combine   information   from   individual   stations,   geological   properties   of   the   site  
(representing  site  amplification),  and  ground  motion  attenuation  for  the  distance  to  
the  epicenter  of  the  causative  fault.    The  idea  is  to  produce  realistic  estimates  at  grid  
points   located   far   from   available   data   while   preserving   the   detailed   shaking  
information   existing   for   regions   where   there   are   stations   nearby.   Figure   2.14  
illustrate  an  example  ShakeMap  for  the  magnitude  7.3  earthquake  in  Southern  Peru.    

 
Figure  2.14.  ShakeMap  for  the  September  25,  2013  magnitude  7.3  in  Southern  Peru.    

16
Chapter  3  

2.4. Ground  Motion  Fields  


Whereas   ShakeMaps   provide   a   rapid,   illustrative   description   of   ground   shaking   after  
an   event   has   occurred,   risk   modelling   requires   the   realistic   prediction   of   ground  
shaking  for  future  possible  events.  The  maps  of  ground  shaking  for  a  future  scenario  
earthquake   (of   which   a   set   are   produced   to   model   the   uncertainty   in   the   ground  
motion)  are  termed  “ground  motion  fields”  herein.    

A   value   of   a   ground   motion   parameter   (e.g.   PGA)   is   calculated   for   a   set   of   sites   using  
a  GMPE  or  IPE  for  a  given  rupture  model  (with  given  geometry  and  magnitude).  The  
median   ground   motion   can   be   calculated   at   each   location,   leading   to   a   median  
ground  motion  field  as  depicted  in  Figure  2.15.    

−78˚ −76˚ −74˚ −72˚ −70˚

−6˚ −6˚

−8˚ −8˚

−10˚ −10˚

−12˚ −12˚

−14˚ −14˚

−16˚ −16˚

−18˚ −18˚
−78˚ −76˚ −74˚ −72˚ −70˚

PGA

0.0 0.1 0.2 0.3 0.4  


Figure  2.15.  Example  of  a  median  ground  motion  field,  considering  the  2007  Pisco  earthquake  in  Peru.    

However,  as  discussed  previously,  there  is  a  large  variability  in  the  level  of  shaking  
at  a  given  site  when  this  is  predicted  with  an  empirical  equation.  In  order  to  account  
for   this   variability,   a   number   of   ground   motion   fields   should   be   generated,   so   that   at  
a  given  site  the  full  range  of  possible  levels  of  ground  motion  are  modelled.  Hence,  
once  the  median  (or  logarthimic  mean)  ground  motion  at  each  site  is  estimated,  an  
estimate  of  the  number  of  logarithmic  standard  deviations  above  or  below  the  mean  
(the   residual)   needs   to   be   made   by   randomly   drawing   from   a   normal   distribution.  
The  proximity  of  the  sites  can  be  accounted  to  consider  that  sites  close  to  each  other  
will   have   similar   residuals   (known   as   spatial   correlation).   Examples   of   ground  
motion   fields   with   and   without   spatial   correlation   are   illustrated   in   Figure   2.16b.  
There   are   a   number   of   spatial   correlation   models,   but   the   most   recent   is   that   of  

17
Basic concepts of seismic hazard

Jayaram   and   Baker   (2009),   which   is   based   on   spectral   acceleration.   Spatial  


correlation   models   for   macroseismic   intensity   have   not   been   published   in   the   past  
but  are  the  focus  of  recent  research.  A  discussion  of  the  modelling  of  uncertainty  in  
ground   motion   for   the   purposes   of   loss   modelling   is   presented   in   Crowley   et   al.  
(2008).  

−78˚ −76˚ −74˚ −72˚ −70˚

−6˚ −6˚

−8˚ −8˚

−10˚ −10˚

−12˚ −12˚

−14˚ −14˚

−16˚ −16˚

−18˚ −18˚
−78˚ −76˚ −74˚ −72˚ −70˚

PGA

0.0 0.1 0.2 0.3 0.4


                                             
Figure  2.16.  Ground  motion  field  without  spatial  correlation  (left)  and  with  spatial  correlation  (right),  
for  the  2007  Pisco  earthquake  in  Peru.  

2.5. Probabilistic  Seismic  Hazard  Assessment  


From   a   probabilistic   seismic   hazard   assessment   (PSHA)   it   is   possible   to   derive  
hazard   curves   and   maps.   A   hazard   curve   presents,   at   a   given   location,   the  
probabilities   of   exceedance   within   a   given   time   period   of   a   range   of   levels   of   ground  
shaking.   A   hazard   map   presents   the   expected   level   of   ground   shaking   for   a   given  
probability   of   exceedance   within   a   certain   time   interval   throughout   the   region   of  
interest.   Figure   2.17   depicts   a   hazard   map   for   a   probability   of   exceedance   of   10%   in  
50  years  for  South  America.    

18
Chapter  3  

 
Figure  2.17.  Seismic  hazard  map  in  terms  of  peak  ground  acceleration  with  a  10%  probability  of  
exceedance  in  50  years  for  South  America.  

The  four  main  steps  to  a  PSHA  are  shown  in  Figure  2.18.  The  first  step  involves  the  
definition  of  seismic  sources  that  are  capable  of  generating  future  earthquakes.  It  is  
generally  assumed  that  any  location  within  a  source  zone  has  an  equal  likelihood  of  
generating   an   event   (i.e.   the   seismicity   is   uniform).   Tectonics,   observed   seismicity  
and  a  degree  of  judgement  are  all  called  upon  in  the  determination  of  source  zones.  
Once   the   source   zones   have   been   defined,   the   temporal   element   of   seismicity   is  
provided   through   recurrence   relationships.   Within   each   source   zone,   a   recurrence  
relationship  can  be  derived  through  the  extraction  of  the  number  of  earthquakes  of  
different   magnitude   from   the   earthquake   catalogue.   A   recurrence   relationship  
specifies   the   average   rate   at   which   an   earthquake   of   some   magnitude   will   be  
exceeded,  the  first  being  proposed  by  Gutenberg  and  Richter  (1994):  

log 𝑁! = 𝑎 − 𝑏𝑀  

where   Nm   is   the   mean   annual   rate   of   exceedance   of   the   magnitude   M,   10!   is   the  
mean  yearly  rate  of  earthquakes  of  magnitude  greater  than  or  equal  to  zero,  and  b  
describes  the  relative  likelihood  of  large  and  small  earthquakes.  The  a  and  b  values  
can   be   obtained   through   regression   analysis   of   an   earthquake   catalogue,   which   is  
pre-­‐processed   to   remove   fore-­‐   and   aftershocks.   The   third   step   involves   the  
definition  of  the  ground  motions  at  the  site,  using  a  GMPE  as  discussed  previously.  
The   fourth   step   involves   the   calculation   of   a   hazard   curve   at   a   given   site   by  
integrating   the   exceedances   of   ground   motions   caused   by   earthquakes   of   different  
sizes   (magnitudes)   occurring   at   varying   locations   in   each   source   zone.   The   actual  
details   of   the   hazard   calculations   are   not   provided   herein   as   this   goes   beyond   the  

19
Basic concepts of seismic hazard

scope   of   this   course.   Further   details   can   be   found   in   Reiter   (1990)   and   McGuire  
(2004).  

 
Figure  2.18.  The  four-­‐step  process  to  a  typical  PSHA  (adapted  from  Reiter,  1990)  leading  to  the  
calculation  of  a  seismic  hazard  curve  (bottom  right)  

Other   intermediate   products   that   can   be   obtained   from   a   PSHA   include   stochastic  
event   sets   and   associated   ground   motion   fields.   Given   a   seismic   source   model,   one  
can   generate   a   number   of   seismicity   histories   (i.e.   randomly   generated   earthquake  
catalogue),   each   one   representing   a   possible   realisation   of   the   seismicity   originating  
from   a   seismic   source   model   during   a   given   time   span.   The   ensemble   of   these  
seismicity   histories   is   called   a   Stochastic   Event   Set.   Figure   2.19   shows   the   events  
within  a  stochastic  event  set  for  Italy  made  up  of  two  different  seismicity  histories  of  
50  years  each.  

   
Figure  2.19.  Two  different  50  year  seismicity  histories  that  make  up  a  stochastic  event  set.  

20
Chapter  3  
Basic  concepts  of  exposure  
modelling  
Exposure  data  as  defined  herein  refers  to  the  elements  that  are  exposed  to  seismic  
hazard,  such  as  people,  buildings  and  infrastructure.    

3.1. Global  exposure  modelling  


There   are   global   datasets   of   gridded   population   with   resolution   of   up   to   30   arc  
seconds  (approximately  1km  square  at  the  ecuator),  some  of  which  are  proprietary  
(e.g.   LandScanTM)   and   others   that   are   open   (e.g.   Gridded   Population   of   the   World,  
version   3,   see   Figure   3.1,   and   GRUMP   v1).   GRUMP   (Global   Urban   and   Rural   Mapping  
Project)   also   identifies   whether   a   given   grid   cell   is   urban   or   rural.   A   review   of   global  
datasets  for  population  has  been  carried  out  by  Salvatore  et  al.  (2005),  but  some  key  
information  on  three  main  datasets  is  provided  below.    

 
Figure  3.1.  Spatial  distribution  of  population  count  according  to  LandScan  2012.  
Basic concepts of seismic risk

LandScanTM   uses   a   large   number   of   data   sources   such   as   maps   of   major   roads   and  
rail   networks,   drainage   networks,   utility   systems,   elevation   contours,   coastlines,  
international   boundaries   and   populated   places,   land   cover   data,   high   resolution  
aerial   photography   and   satellite   imagery,   and   the   US   Bureau   of   Census   data   to  
produce   a   30   arc   second   grid   of   population.   It   is   noted   that   the   population   is   an  
ambient   population,   as   opposed   to   a   residential   population,   and   thus   takes   into  
account  the  movements  of  individuals  through  a  given  area  (or  cell).    

GPW   version   3   is   a   2.5   arc   minutes   (approximately   5km   square   at   the   equator)  
spatially   disaggregated   population   layer   based   on   national   statistical   office  
estimates   of   population,   where   the   population   is   distributed   uniformly   across   the  
grid   within   a   given   administrative   unit   for   which   the   census   data   is   available.  
Population  data  estimates  are  provided  for  1990,  1995,  and  2000,  and  projected  to  
2005,  2010,  and  2015.    

GRUMP   v1   built   upon   GPW   to   construct   a   geo-­‐referenced   framework   of   urban   and  


rural   areas   by   combining   census   data   with   satellite   data.   However,   GRUMPv1  
actually   comprises   three   main   data   products:   1)   a   higher   resolution   gridded  
population  data  product  at  30  arc-­‐seconds,  or  ~1km  at  the  equator,  for  1990,  1995,  
and  2000;  2)  urban  extents  dataset  delineating  urban  areas  based  on  NOAA’s  night-­‐
time   lights   data   set   and   buffered   settlement   centroids   (where   night   lights   are   not  
sufficiently  bright);  3)  a  points  dataset  of  all  urban  areas  with  populations  of  greater  
than   1,000   persons.   These   two   datasets   report   residential   as   opposed   to   ambient  
population.    

The  only  currently  available  dataset  related  to  building  inventory  on  a  global  scale  is  
the   USGS   PAGER   database   v1.4   (Jaiswal   and   Wald,   2008)   and   the   Global   Exposure  
Database  of  the  Global  Earthquake  Model.  Regarding  the  former  dataset,  a  building  
taxonomy   (describing   typologies   of   buildings   and   their   characteristics)   was  
developed   and   datasets   with   residential   housing   building   stock   distributions  
available   for   individual   countries   (from   a   variety   of   sources   such   as   UN-­‐HABITAT,  
housing   census,   published   literature,   expert   opinion)   were   mapped   to   these  
construction  types.  Where  no  data  was  available  for  a  given  country,  the  distribution  
of   housing   stock   from   selected   neighbouring   countries   was   used.   Figure   3.2  
represents  the  various  sources  of  data  used  for  each  country  in  the  World  used  by  
PAGER.    

22
Chapter  3  

Figure  3.2.  PAGER  global  building  inventory  data  sources  

The  distributions  of  occupants  in  building  types  were  divided  between  urban,  rural,  
residential   and   non-­‐residential.   Figure   3.3   provides   an   extract   of   the   database   for  
Chile.   Algorithms   to   further   distribute   these   occupants   during   the   day,   night   and  
transit   times   were   also   developed.   These   distributions   can   be   used   in   conjunction  
with   the   aforementioned   global   population   databases   to   estimate   the   occupants  
within  different  building  typologies  at  different  times  during  the  day.    

 
Figure  3.3.  Building  distribution  per  type  of  area/use  for  Chile.  

To   improve   on   the   aforementioned   PAGER   database,   the   Global   Earthquake   Model  


has  funded  the  development  of  the  Global  Exposure  Database,  that  combines  many  

23
Basic concepts of seismic risk

sources   of   data   to   produce   a   gridded   distribution   of   buildings   (counts,   area,  


replacement   cost)   and   their   occupants,   distributed   according   to   their   structural  
characteristics  (following  the  GEM  building  taxonomy,  as  described  in  the  following  
section).   Huyck   et   al.   (2011)   provide   a   detailed   description   of   additional   datasets  
that   are   available   around   the   world   (at   global   and   regional   scales)   to   support   the  
modelling   of   global   exposure   data,   such   as   administrative   areas   (boundaries   of  
countries,  regions,  provinces,  municipalities  etc.),  population  datasets,  building  data  
(from  census  data  and  surveys),  land  use  and  land  cover  datasets.    

3.2. Regional  exposure  modelling  


The  development  of  an  exposure  model  capable  of  providing  information  about  the  
location,  value  and  vulnerability  classification  of  the  exposed  elements  at  a  national  
or   regional   scale   might   be   a   very   challenging   task,   often   only   achievable   through   the  
use   of   Census   data   (e.g.   Erdik   et   al.   2003   (Turkey);   Silva   et   al.   2014a   (Portugal);  
Crowley   et   al.   2008   (Italy)).   The   previously   described   approach   can   also   be  
employed,  but  the  reliability  of  the  final  exposure  model  might  not  be  sufficient  for  
regional   or   local   risk   analysis.   These   datasets   may   contain   information   about   the  
number   of   dwellings   and/or   building,   material   of   the   walls/roof/floor,   number   of  
stories,   type   of   dwelling   and   date   of   construction.   With   this   information,   it   is  
possible   to   associate   each   asset   with   a   vulnerability   class,   and   calculate   a   spatial  
distribution   of   these   categorized   assets   throughout   the   region   of   interest.   Figure   3.4  
illustrates   a   regional   model   that   has   been   developed   by   the   Global   Earthquake  
Model  (SARA  project)  for  the  Andean  countries  following  this  approach.  

 
Figure  3.4.  Exposure  modelling  for  the  Andean  countries.  

24
Chapter  3  

3.3. The  importance  of  Building  Taxonomies  


A   taxonomy   is   a   classification   scheme,   and   a   number   of   schemes   relevant   to   the  
classification   of   buildings   for   seismic   risk   assessment   have   been   proposed   in   the  
past.   These   taxonomies   have   in   general   been   developed   to   describe   and   classify  
building  structures  in  terms  of  seismic  resistance  and  response.  They  have  all  been  
developed   since   1985   and   can   be   expected   to   have   improved   upon   earlier   similar  
taxonomies.   The   primary   requirements   of   the   proposed   structural   taxonomy   are  
listed  below.    

The   criteria   by   which   the   existing   taxonomies   are   assessed   are   listed   below:  
Distinguishes   differences   in   seismic   performance.   The   taxonomy   distinguishes  
earthquake-­‐resistant   versions   of   structural   systems   from   non-­‐earthquake   resistant  
version,   including   the   "   before"   and   "after"   states   of   common   seismic   retrofits   and  
between  ductile  and  non-­‐ductile  systems.    

 Observable.   Two   people   examining   the   same   structural   system   in   the   field   or   using  
data  obtained  from  the  field  should  independently  assign  the  same  taxonomic  group  
based  solely  on  the  text  definition  of  the  taxonomic  group.  

Complete.  The  taxonomy  must  include  all  engineering  features  relevant  to  the  global  
seismic   performance   of   a   building   structure.   As   mentioned   above   it   is   recognized  
that   there   will   be   a   need   for   additional   taxonomies   to   capture   all   aspects   of   the  
seismic   performance   and   losses   for   an   entire   building,   including   building  
dimensions   and   nonstructural   components.   The   structural   taxonomy   must   contain  
sufficient  attributes  to  meet  the  needs  of  the  end  users  of  GEM.  

Simple   and   collapsible.   The   taxonomy   should   have   as   few   groups   as   possible,   while  
still   meeting   the   other   requirements.   It   is   also   desirable   to   define   common  
combinations   and   relative   quantities   of   structural   systems   so   that   fragility   or  
vulnerability  functions  can  be  created  by  aggregating  the  fragilities  or  vulnerabilities  
of   detailed   components,   while   still   distinguishing   say,   differences   in   ductility,   design  
or  retrofit  alternatives.  A  taxonomy  is  judged  to  be  collapsible  if  taxonomic  groups  
can   be   combined   and   the   resulting   combinations   still   distinguish   differences   in  
seismic  performance.  

Nearly   exhaustive.   Within   practical   limits   almost   every   structural   system   can   be  
sensibly  assigned  to  a  taxonomic  group.  

25
Basic concepts of seismic risk

Familiar  to  engineering  practitioners  and  architects.  It  is  desirable  that  engineers  
and   architects   be   familiar   with   the   taxonomic   system,   particularly   to   readily   and  
accurately   identify   structural   attributes.   If   the   new   taxonomic   system   corresponds  
readily   to   an   existing   taxonomic   system,   it   can   give   users   access   to   existing   data.  
Engineers  and  architects  should  be  familiar  with  the  nomenclature  to  be  defined  to  
avoid  ambiguity.  

Currently   treats   non-­‐buildings.   Built   forms   other   than   buildings   need   to   be  


included   in   the   taxonomy   sometime   in   the   future.   These   include   constructions   like  
dams,  bridges  and  tunnels.  

Extensible  to  other  hazards.  In  the  future  parts  of  the  GEM  model  may  be  used  by  
other   groups   working   on   other   natural   hazards   such   as   floods,   hurricanes   and  
volcanic  eruptions.  

User-­‐friendly.  The  taxonomy  should  be  straightforward,  intuitive,  and  as  easy  to  use  
as  possible  by  both  those  collecting  data,  those  arranging  for  its  analysis  and  the  end  
users.  

International   in   scope.   As   far   as   possible   the   taxonomy   should   be   appropriate   for  


any  region  of  the  world.  It  should  not  privilege  any  one  region  but  be  technically  and  
culturally  acceptable  to  all  regions.  

3.4. GEM  Basic  Building  Taxonomy  


The  GEM  Basic  Building  Taxonomy  relies  on  8  key  attributes,  as  reported  in  the  table  
below.   See   Brzev   et   al.   2013   for   the   full   report   on   the   taxonomy,   and   all   of   the  
reference  tables.  

Table  3.1  GEM  Basic  Building  Taxonomy  summary  table    


Attribute   Attribute  levels   Example  

Material  of  the  Lateral   Material  type  (Level  1)  


Load-­‐Resisting   Material  technology  (Level  2)   Steel  
System   Material  properties  (Level  3)  
Type  of  lateral  load-­‐resisting  system  (Level  1)  
Lateral  Load-­‐Resisting    
Braced  frame  
System   Ductility  (Level  2)  
 

Roof  material  (Level  1)   Masonry  


Roof    
Roof  type  (Level  2)  
 
Floor  material  (Level  1)  
Floor   Concrete  
Floor  type  (Level  2)  

26

 
Chapter  3  

Height   Number  of  stories  above  ground   4  


Date  of  Construction   Construction  completed  (year)   1925  
Type  of  irregularity  (Level  1)  
Irregularity   Re-­‐entrant  corner  
Irregularity  description  (Level  2)  
Building  occupancy  class  -­‐  top  level  (Level  1)  
Occupancy   Residential  
Building  occupancy  class  -­‐  detail  (Level  2)  
 

27
Chapter  4  
Seismic  vulnerability  assessment    
4.1. Main  outputs  in  vulnerability  assessment  
Physical   vulnerability   is   defined   herein   as   the   probability   distribution   of   loss   (or   loss  
ratio)  due  to  structural  or  non-­‐structural  damage,  given  an  intensity  measure  level  
(see   Figure   4.1).   Loss   might   refer   to   repair   costs   or   loss   of   life;   loss   ratio   might   refer  
to   the   ratio   of   cost   of   repair   to   cost   of   replacement,   or   the   ratio   of   number   of  
fatalities   to   the   number   of   occupants.   These   vulnerability   functions   can   be   derived  
directly,   usually   through   empirical   methods   where   the   losses   from   past   events   at  
given   locations   are   related   to   the   levels   of   intensity   of   ground   motion   at   those  
locations,   or   they   can   be   derived   indirectly   by   combining   fragility   functions   and  
consequence  functions,  as  described  in  the  following  sections.    

0.7

0.6

0.5
Loss ratio (%)

0.4

0.3

0.2

0.1

0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
Spectral acceleration at T=0.30s (g)  
Figure  4.1.  Discrete  vulnerability  function  
Chapter  4  

Fragility  functions  describe  the  probability  of  exceeding  a  set  of  limit  states,  given  an  
intensity  measure  level;  limit  states  describe  the  limits  to  performance  levels,  such  
as   damage   or   injury   levels.   Fragility   functions   can   be   derived   by   expert-­‐opinion,  
empirically   (using   observed   data),   or   analytically,   by   explicitly   modelling   the  
behaviour   of   a   given   asset   typology   when   subjected   to   increasing   levels   of   ground  
motion.  The  relation  between  probability  and  ground  shaking  can  be  modelled  by  a  
cumulative  lognormal  function  (Figure  4.2),  or  in  a  discrete  manner,  in  which  each  
intensity   measure   level   is   simply   associated   with   a   probability   of   exceedance   per  
damage  state  (Figure  4.3).  

 
Figure  4.2.  Continuous  fragility  functions  

1
0.9
0.8
Probability of exceedance

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
5 6 7 8 9 10 11
MMI

Negligible Moderate Substantial Very Heavy Destruction


 
Figure  4.3.  Discrete  fragility  functions  

Consequence  functions  are  used  to  derive  the  probability  distribution  of  loss,  given  a  
performance  level,  and  are  generally  derived  empirically  or  through  expert  opinion.  
Figure  4.4  illustrates  a  consequence  model  for  buildings  in  Greece.  

29
Basic concepts of seismic risk

1"

0.8"

Damage"ra>o"
0.6"

0.4"

0.2"

0"

Slight"

Moderate"

Collapse"
Heavy"

Very"heavy"
 
Figure  4.4.  Consequence  model  showing  mean  ratios  of  cost  of  repair  to  cost  of  replacement,  and  bars  
illustrating  the  uncertainty  in  these  ratios  (adapted  from  Kappos  et  al.  2006).  

4.2. Expert  Opinion  Vulnerability  


Expert   opinion   vulnerability   assessment   is   needed   in   some   situations,   for   example  
due  to  the  following  reasons  (Jaiswal,  personal  communication):  

1. Observed  damage  and  loss  data  exist  only  at  places  where  large  earthquakes  
have  occurred  in  the  past;  

2. Structural  modelling  and  analysis  techniques  for  estimating  damage  and  loss  
are   not   always   suitable   to   model   structural   collapse   behaviour   of   a   large  
number  of  structure  types  in  different  parts  of  the  world;  

3. It  is  not  practical  to  perform  hundreds  of  experiments/shake-­‐table  tests  on  
real-­‐life  structures  as  they  are  built  today;  

4. Structural   design   and   construction   practice   may   differ   substantially   over  


time  and  space,  and  it  is  practically  difficult  to  fully  incorporate  such  changes  
appropriately  within  available  analytical  models;  

Vulnerability   functions   were   produced   using   expert   opinion   as   part   of   the   ATC-­‐13  
project   (ATC,   1985).   More   than   50   senior   earthquake   engineering   experts   were  
asked  to  provide  low,  best  and  high  estimates  of  the  loss  ratios  (the  ratio  of  repair  
cost   to   replacement   cost,   expressed   as   a   percentage,   which   they   termed   “damage  
factor”)   for   Modified   Mercalli   Intensities   (MMI)   from   VI   to   XII   for   40   different  
building  classes.  The  low  and  high   loss  ratio  estimates  provided  by  the  experts  were  
defined   as   the   90%   probability   bounds   of   a   lognormal   distribution,   whilst   the   best  
estimate   was   taken   as   the   median   loss   ratio   (Figure   4.5).   Weighted   means   of   the  
experts’  estimates,  based  on  the  experience  and  confidence  levels  of  the  experts  for  

30
Chapter  4  

each   building   class,   were   included   in   the   averaging   process,   as   described   in  


Appendix  G  of  ATC-­‐13  (ATC,  1985).    

5% 5%

ML MB MH
4.00 10.60 18.20

Figure  4.5    Example  of  a  lognormal  distribution  of  the  estimated  loss  ratio  for  a  given  macroseismic  
intensity  showing  the  mean  low  (ML),  mean  best  (MB)  and  mean  high  (MH)  estimates  of  the  damage  
factor  (adapted  from  McCormack  and  Rad,  1997).  

For   each   intensity   measure   level   in   the   ATC-­‐13   vulnerability   model,   a   mean   loss  
ratio   (so-­‐called   damage   factor)   and   corresponding   coefficient   of   variation   is   defined.  
A   lognormal   distribution   of   the   loss   ratio   for   each   intensity   measure   level   is  
assumed.   Figure   4.6   presents   the   ATC-­‐13   vulnerability   functions   (mean   loss   ratio  
only)  for  moment-­‐resisting  concrete  frame  buildings.  

The   Global   Earthquake   Model   funded   an   initiative   that   was   carried   out   by   the  
physical   vulnerability   consortium   to   build   upon   the   ATC-­‐13   project   to   develop  
expert  opinion  fragility/vulnerability  assessment  for  a  number  of  building  types  for  
which  empirical  or  analytical  functions  are  not  readily  available.  A  new  methodology  
for   weighting   the   experts’   opinions   (Cooke’s   method)   was   implemented.   More  
information  on  the  results  of  this  project  are  available  on  GEM  Nexus.    

 
Figure  4.6  ATC-­‐13  vulnerability  functions  (mean  loss  ratio  versus  intensity  measure  level,  IML)  for  
reinforced  concrete  moment-­‐resisting  frame  buildings  

31
Basic concepts of seismic risk

It   might   be   argued   that   expert   opinion   is   close   to   empirical   fragility/vulnerability  


assessment  as  the  experts  are  likely  to  be  basing  their  opinion  on  their  experience  in  
observing  damage  and  loss  in  the  field.    

4.3. Empirical  Vulnerability  


Empirical   vulnerability   uses   observations   of   loss   and   relates   these   to   the   observed  
ground  shaking  to  produce  vulnerability  functions.    

An   empirical   global   vulnerability   model   capable   of   predicting   human   losses  


(fatalities)  has  been  developed  by  Jaiswal  et  al.  (2008),  to  rapidly  and  approximately  
ascertain  the  impact  in  terms  of  fatalities  immediately  following  a  large  earthquake  
anywhere   in   the   world.   In   order   to   develop   this   model,   earthquake   mortality   rates  
(attributed   to   damage   from   ground   shaking)   for   more   than   4,800   worldwide  
earthquakes   that   occurred   since   1973   were   studied   and   a   set   of   vulnerability  
functions   for   specific   countries   or   regions   were   built.   This   earthquake   fatality   rate  
can   be   defined   as   the   ratio   of   the   total   number   of   shaking-­‐related   fatalities   to   the  
total   population   exposed   at   a   given   shaking   intensity,   using   Modified   Mercalli   as   the  
intensity  measure  type.    

The   PAGER   collection   of   vulnerability   functions   that   return   a   fatality   rate   for   a   given  
intensity   measure   level,   can   be   expressed   in   terms   of   a   two-­‐parameter   lognormal  
cumulative   distribution.   For   countries   where   at   least   4   fatal   seismic   events   have  
occurred   (earthquakes   where   casualties   occurred)   a   specific   vulnerability   function  
was   computed.   However,   only   a   few   countries   have   experienced   >4   fatal  
earthquakes  since  1973  and  therefore,  it  was  necessary  to  aggregate  some  countries  
in   regions   in   order   to   have   enough   fatal   earthquakes   for   this   catalogue   period.   For  
this   reason,   a   new   global   regionalization   scheme   was   proposed   based   on   the  
likelihood   of   suffering   similar   losses   in   future   earthquakes.   This   spatial   organization  
was   done   based   primarily   on   geography,   building   inventory   and   social-­‐economic  
similarities.  Figure  4.7  presents  the  regionalization  proposed  by  the  aforementioned  
methodology,   where   countries   with   the   same   colour   all   use   the   same   empirical  
vulnerability   function.   According   to   the   proposed   regionalization,   the   vulnerability  
model  is  composed  of  27  functions  for  specific  countries  and  23  functions  for  groups  
of   countries.   These   expressions,   which   give   fatality   rate   (ν)   (i.e.   ratio   of   number   of  

32
Chapter  4  

fatalities  to  exposed  population)  as  a  function  of  shaking  intensity  (S),  are  expressed  
as  a  lognormal  distribution  function  as  follows:  

1 𝐼
𝜈 𝑆 =𝜙 𝑙𝑛  
𝛽 𝜃

where   φ   is   the   standard   normal   cumulative   distribution   function,   I   is   the   discrete  


value   of   the   intensity   measure   level   and   β   and   θ   are   the   parameters   of   the  
distribution.  This  vulnerability  model  is  bounded  between  intensity  V  to  IX  (where  
populations   exposed   to   shaking   intensity   IX   and   above   are   collated   at   shaking  
intensity  IX).  The  two  parameters  of  the  distribution  were  computed  in  a  way  that  
the   residual   error   would   be   minimized   for   larger   and   smaller   earthquakes  
simultaneously.    

 
Figure  4.7  Regionalisation  scheme  for  PAGER  empirical  casualty  vulnerability  model.  

Jaiswal   and   Wald   (2011)   extended   the   aforementioned   methodology   to   develop  


empirical   vulnerability   functions   in   terms   of   economic   loss.   In   order   to   estimate  
economic   losses,   an   assessment   of   the   economic   exposure   at   various   levels   of  
shaking   intensity   was   required.   The   economic   value   of   all   the   physical   assets  
exposed  at  different  locations  in  a  given  area  is  generally  not  known  and  extremely  
difficult  to  compile  at   a  global  scale.  In  the  absence   of  such  a   dataset,   the  total  Gross  
Domestic   Product   (GDP)   exposed   at   each   shaking   intensity   was   estimated   by  
multiplying  the  per-­‐capita  GDP  of  the  country  by  the  total  population  exposed  at  that  

33
Basic concepts of seismic risk

shaking  intensity  level.  The  total  GDP  estimated  at  each  intensity  was  then  scaled  by  
an   exposure   correction   factor,   which   is   a   multiplying   factor   calculated   for   each  
country   and   which   aims   to account   for   the   disparity   between   wealth   and/or  
economic   assets   to   the   annual   GDP.   The   economic   exposure   obtained   using   this  
procedure   is   thus   a   proxy   estimate   for   the   economic   value   of   the   actual   inventory  
that   is   exposed   to   the   earthquake.   The   economic   loss   ratio,   defined   in   terms   of   a  
country-­‐specific  lognormal  cumulative  distribution  function  of  shaking  intensity,  has  
been  derived  and  calibrated  against  the  losses  from  past  earthquakes.  Economic  loss  
data   available   for   global   earthquakes   from   1980   to   2007   from   the   MunichRe  
NatCatSERVICE   databasehas   been   used   for   this   purpose.   The   same   regionalization  
scheme   as   proposed   for   the   fatality   model   (Figure   4.7)   is   used   with   the   economic  
vulnerability  model.  

4.4. Fragility  Assessment  

4.4.1. Damage  States  


The   two   previous   sections   dealt   with   the   direct   estimation   of   vulnerability   (i.e.  
distribution  of  loss  ratio  given  intensity),  whereas  as  described  in  the  introduction,  
vulnerability  can  be  estimated  through  damage,  by  first  estimating  the  distribution  
of   damage   (using   fragility   functions)   and   then   relating   that   damage   distribution   to  
loss  (using  consequence  functions).  

There   are   many   structural   damage   scales   that   consist   of   discrete   damage   state  
descriptions   that   are   meant   to   represent   different   levels   of   building   performance,  
from   no   damage   to   collapse.   Damage   scales   can   be   used   to   assign   damage   to  
buildings  for  post-­‐earthquake  damage  and  safety  assessment,  or  for  the  definition  of  
performance   of   buildings   in   analytical   vulnerability   assessment.   Existing   damage  
scales   thus   vary   greatly   in   the   number   of   damage   scales   adopted   (seen   to   vary  
between  3  and  12  in  the  literature),  details  of  the  damage  state  descriptions,  damage  
criteria   used   and   the   amount   and   type   of   information   needed   for   the   damage  
description.   An   example   damage   scale   for   Reinforced   Concrete   Moment   Resisting  
Frames  (so-­‐called  C1)  according  to  HAZUS  (FEMA,  2003)  is  provided  in  Table  4.1.    

34
Chapter  4  

Table  4.1  Damage  scale  from  HAZUS  (FEMA,  2003)  for  Reinforced  Concrete  Moment  Resisting  Frames  
(C1)  

Damage  State   Damage  Description  

Flexural  or  shear  type  hairline  cracks  in  some  beams  and  
Slight  Structural  Damage  
columns  near  joints  or  within  joints.  
Most  beams  and  columns  exhibit  hairline  cracks.  In  ductile  
frames  some  of  the  frame  elements  have  reached  yield  
Moderate  Structural  Damage   capacity  indicated  by  larger  flexural  cracks  and  some  
concrete  spalling.  Non-­‐ductile  frames  may  exhibit  larger  
shear  cracks  and  spalling.  
Some  of  the  frame  elements  have  reached  their  ultimate  
capacity  indicated  in  ductile  frames  by  large  flexural  
cracks,  spalled  concrete  and  buckled  main  reinforcement;  
Extensive  Structural  Damage   non-­‐ductile  frame  elements  may  have  suffered  shear  
failures  or  bond  failures  at  reinforcement  splices,  or  
broken  ties  or  buckled  main  reinforcement  in  columns  
which  may  result  in  partial  collapse.  
Structure  is  collapsed  or  in  imminent  danger  of  collapse  
due  to  brittle  failure  of  non-­‐ductile  frame  elements  or  loss  
of  frame  stability.  Approximately  13%(low-­‐rise),  
Complete  Structural  Damage  
10%(mid-­‐rise)  or  5%(high-­‐rise)  of  the  total  area  of  C1  
buildings  with  Complete  damage  is  expected  to  be  
collapsed.  
 

Hill   and   Rossetto   (2008a)   provide   a   detailed   review   and   comparison   of   damage  
scales  for  different  building  types  (Figure  4.8).    

 
Figure  4.8  Comparison  of  damage  scales  for  reinforced  concrete  with  masonry  infill  (adapted  from  Hill  
and  Rossetto,  2008a)    

4.4.2. Fragility  Functions  and  Damage  Probability  Matrices  


Fragility  functions  can  be  discrete  or  continuous.  A  continuous  fragility  function  was  
shown   previously   in   Figure   4.2.   When   macroseismic   intensity   is   the   intensity  
measure,   the   functions   are   generally   discrete   (due   to   the   integer   scale   of  
macroseismic  intensity),  such  as  the  example  shown  in  Figure  4.9.    

35
Basic concepts of seismic risk

 
Figure  4.9.  Discrete  fragility  functions;  the  distribution  of  damage  at  MMI  =  8.5  is  indicated.  

This   fragility   function   can   also   be   represented   in   terms   of   a   damage   probability  


matrix,   where   for   a   given   intensity   measure   level   (e.g.   MMI   =   8)   the   percentage   of  
buildings   within   each   damage   state   is   reported   (obtained   by   probability   of   attaining  
or   exceeding   that   damage   state   minus   the   probability   of   attaining   or   exceedance   the  
next  damage  state).    

Table  4.2  Damage  probability  matrix  (after  Whitman  et  al.  1973)  

Non-­‐ Intensity  of  Earthquake  


Damage   Structural  
Structural  
State   Damage  
Damage   V   VI   VII   VIII   IX  

0   None   None   10.4   -­‐   -­‐   -­‐   -­‐  


1   None   Minor   16.4   0.5   -­‐   -­‐   -­‐  
2   None   Localised   40.0   22.5   -­‐   -­‐   -­‐  
3   Not  noticeable   Widespread   20.0   30.0   2.7   -­‐   -­‐  
4   Minor   Substantial   13.2   47.1   92.3   58.8   14.7  
5   Substantial   Extensive   -­‐   0.2   5.0   41.2   83.0  
6   Major   Nearly  total   -­‐   -­‐   -­‐   -­‐   2.3  
7   Building  condemned   -­‐   -­‐   -­‐   -­‐   -­‐  
8   Collapse   -­‐   -­‐   -­‐   -­‐   -­‐  
 

4.4.3. Expert  Opinion  Fragility  


A  method  for  fragility  assessment  that  can  be  placed  in  the  expert  opinion  category  
is   that   proposed   by   Lagomarsino   and   Giovinazzi   (2006),   which   is   based   on   the   EMS-­‐
98   macroseismic   scale.   The   outputs   of   this   method   include   damage   probability  
matrices,   and   mean   damage   factor   functions,   which   can   then   be   combined   with  
consequence  functions  to  produce  vulnerability  functions.    

36
Chapter  4  

According   to   the   European   Macroseismic   Scale   (Grunthal,   1998),   intensity   can   be  


evaluated   either   by   direct   field   observations   or   by   questionnaires   sent   to   the  
affected   areas,   although   the   latter   is   really   only   suitable   for   lower   intensity   values.  
The  “modal”  intensity  is  assigned  to  a  given  area  based  on  the  observations  for  the  
following  three  categories:  (a)  descriptions  of  people’s  reactions,  (b)  the  movement  
of   objects   and   (c)   the   proportions   of   each   building   category   expected   to   be   within  
each  damage  state.    

Building   damage   begins   at   intensity   grade   V,   whereas   the   degrees   of   intensity   VII,  
VIII  and  IX  on  the  EMS  scale  describe  the  most  important  degrees  of  shaking  in  many  
ways,   from   an   engineering   point,   since   they   are   frequently   encountered   and  
potentially  damaging.  Intensity  is  sometimes  thought  of  as  an  indication  of  building  
damage,   but   this   is   really   a   misperception.   Different   types   of   building   are   treated  
differently   and   levels   of   damage   are   carefully   defined   precisely   to   deconvolute   the  
ground   motion   from   the   response   of   buildings.   Figure   4.10   shows   how   EMS-­‐98  
classifies  the  vulnerability  class  of  different  types  of  building.  Figure  4.11  illustrates  
the  different  grades  of  damage  for  a  given  building  typology.  Buildings  are  assigned  
to   a   vulnerability   class,   and   the   damage   in   buildings   within   a   given   vulnerability  
class   is   observed   and   recorded   (few,   many   or   most   assigned   to   a   given   damage  
grade),   and   then   damage   probability   matrices   are   used   to   assign   a   macroseismic  
intensity  for  the  area.    

37
Basic concepts of seismic risk

 
Figure  4.10.  Relationship  between  building  types  and  vulnerability  classes.  

 
Figure  4.11.  Damage  grade  descriptions  for  a  given  building  type  

38
Chapter  4  

Table  4.3  Damage  Probability  Matrix  (DPM)  for  vulnerability  class  A.  

DAMAGE GRADE
CLASS A
0 1 2 3 4 5
V Few
VI Many Few
INTENSITY
VII Many Few
DEGREE VIII Many Few
IX Many Few
X Many
XI Most
XII All
 

As  described  previously,  the  EMS-­‐98  scale  defines  qualitative  descriptions  of  “Few”,  
“Many”  and  “Most”  for  five  damage  grades  for  levels  of  intensity  ranging  from  V  to  
XII   for   six   different   classes   of   decreasing   vulnerability   (from   A   to   F).   Damage  
matrices   containing   a   qualitative   description   of   the   proportion   of   buildings   that  
belong   to   each   damage   grade   for   various   levels   of   intensity   were   presented  
previously  in  Table  4.3  for  vulnerability  class  A.    

The   problems   related   to   the   vagueness   of   the   matrices   (i.e.   they   are   described  
qualitatively)  and  the  incompleteness  of  the  matrices  (i.e.  the  lack  of  information  for  
all  damage  grades  for  a  given  level  of  intensity)  has  been  tackled  by  Giovinazzi  and  
Lagomarsino   (2004;   2006)   by   applying   Fuzzy   Set   Theory   and   by   assuming   a   beta  
damage   distribution,   respectively.   In   classical   set   theory,   the   membership   of  
elements   in   a   set   is   assessed   in   binary   terms  -­‐   an   element   either   belongs   or   does   not  
belong  to  the  set.  By  contrast,  fuzzy  set  theory  permits  the  gradual  assessment  of  the  
membership   of   elements   in   a   set;   this   is   described   with   the   aid   of   a   membership  
function   valued   in   the   real   unit   interval   [0,  1],   where   a   value   of   0   means   the  
parameter  does  not  belong  to  the  set,  a  value  of  1  indicates  the  degree  of  belonging  
is   almost   sure   whilst   anything   between   0   and   1   means   rare   but   possible.  
Membership  functions  have  been  applied  to  the  percentage  ranges  of  few,  many  and  
most  described  in  EMS-­‐98,  as  shown  in  Figure  4.12.    

39
Basic concepts of seismic risk

 
Figure  4.12  Membership  functions  of  the  percentage  ranges  used  for  the  quantitative  terms  few,  many  
and  most  in  EMS-­‐98  (Giovinazzi  and  Lagomarsino,  2004;  2006).  

Possible  and  plausible  upper  and  lower  bounds  are  assigned  to  each  linguistic  term  
(e.g.  for  many:  the  possible  lower  bound  is  10,  the  plausible  lower  bound  is  20,  the  
possible   upper   bound   is   50,   the   plausible   upper   bound   is   60).   The   discrete   beta  
distribution   is   assumed   for   the   damage   distribution   for   a   given   intensity   level   and  
the  parameters  of  the  distribution  (where  the  mean  relates  the  mean  damage  grade  
varying   between   0   and   5),   which   leads   to   the   matching   of   the   values   of   the   plausible  
and  possible  upper  and  lower  values  of  the  linguistic  terms.  This  leads  to  four  curves  
(mean  damage  grade  versus  EMS-­‐98  intensity)  for  each  vulnerability  class,  as  shown  
in  Figure  4.13.  

 
Figure  4.13  Plausibility  and  possibility  curves  for  vulnerability  class  B  and  C  

To  treat  the  belonging  of  a  given  building  class  to  a  vulnerability  class,  which  is  also  
treated   in   linguistic   terms   in   EMS-­‐98   (see   Figure   4.10),   Fuzzy   Set   Theory   was   also  
used.  Each  vulnerability  class  is  related  to  a  vulnerability  index,  which  depends  on  
the  building  typology,  the  characteristics  of  the  building  stock  (e.g.  number  of  floors,  

40
Chapter  4  

irregularity   etc.)   and   the   regional   construction   practices.   The   vulnerability   index  
membership  function  is  show  in  Figure  4.14.  The  plausible  and  possible  upper  and  
lower  vulnerability  index  bounds  of  each  vulnerability  class  are  then  related  to  the  
four   functions   of   mean   damage   grade   shown   in   Figure   4.13,   and   an   analytical  
function  is  fit  to  the  curves  as  a  function  of  macroseismic  intensity  and  Vulnerability  
Index  as  follows:  

𝐼 + 6.25𝑉! − 13.1
𝜇! = 2.5 1 + 𝑡𝑎𝑛ℎ
2.3

 
Figure  4.14  Vulnerability  index  membership  functions  for  EMS-­‐98  vulnerability  classes  (Giovinazzi  and  
Lagomarsino,  2006)  

4.4.4. Empirical  Fragility  


Empirical   fragility   assessment   requires   the   observation   of   damage   to   different  
building   typologies   under   different   levels   of   ground   shaking.   The   method   relies   on  
the   collection   of   damage   data   for   different   building   types   following   an   earthquake.  
This   is   often   done   for   a   variety   of   reasons:   rapid   structure   evaluations   are   carried  
out   to   ascertain   the   safety   of   the   buildings   in   the   immediate   aftermath   of   an  
earthquake;   detailed   damage   evaluations   are   carried   out   in   certain   regions,   for   a  
subset   of   the   total   number   of   buildings,   perhaps   following   a   request   from   the  
building   owner;   detailed   post-­‐earthquake   surveys   are   undertaken   by  
reconnaissance   teams   on   small   samples   of   the   buildings.   The   ideal   source   of   data  
should   include   all   of   the   buildings   in   a   given   area   (including   the   non-­‐damaged  
buildings),  each  building  should  be  georeferenced,  and  the  data  should  be  collected  
by   structural   engineers   in   areas   close   to   ground   motion   recording   stations.   Such  
data   is,   however,   hard   to   come   by   and   often   assumptions   are   needed   (e.g.   to  

41
Basic concepts of seismic risk

estimate   the   number   of   undamaged   buildings,   to   estimate   the   ground   shaking   levels  
to  which  the  buildings  were  subjected).    

The   choice   of   the   damage   scale   in   empirical   fragility   functions   is   often   limited   to   the  
damage  scale  used  in  the  post-­‐earthquake  damage  survey,  though  when  a  number  of  
different   surveys   need   to   be   combined   (perhaps   from   different   events   in   different  
countries,   or   from   surveys   using   different   data   collection   forms),   then  
harmonisation  to  a  common  damage  scale  is  necessary  and  studies  such  as  the  one  
of  Hill  and  Rossetto  (2008a)  can  be  particularly  useful.    

The  choice  of  the  building  taxonomy  can  also  be  limited  by  the  one  used  on  the  post-­‐
earthquake   damage   survey,   though   as   for   damage   scales,   harmonisation   to   a  
common  taxonomy  might  be  necessary  when  different  surveys  are  combined  in  the  
generation  of  empirical  fragility  functions.  

The   geographical   unit   for   which   damage   statistics   are   developed   should   be   the  
smallest  area  possible  (with  uniform  soil  conditions)  that  will  provide  a  statistically  
representative   sample.   When   the   data   is   not   available   on   a   building-­‐by-­‐building  
basis,  this  choice  of  geographical  unit  is  imposed  by  the  resolution  of  the  data,  and  
might   refer   to   an   administrative   area.   The   minimum   number   of   observations   per  
dataset  and  the  minimum  number  of  data  points  and  their  distribution  with  respect  
to  the   intensity   measure   levels  is  an   open  issue  that   has  been  investigated   by  GEM’s  
Global  Vulnerability  Consortium  (Rossetto  et  al.  2015).  Fragility  functions  have  been  
developed  in  the  past  with  as  few  as  3  data  points.  

The   levels   of   ground   motion   to   which   the   damaged   buildings   were   subjected   is  
usually  the  most  difficult  data  to  obtain,  due  to  a  lack  of  observations/recordings  at  
the   locations   where   the   buildings   were   damaged.   Recent   post-­‐earthquake   damage  
surveys   are   concentrating   their   reconnaissance   efforts   close   to   ground   motion  
recordings,   but   this   does   not   completely   remove   the   uncertainty   in   the   ground  
motion  (as  local  site  conditions  can  vary  the  ground  shaking  even  over  distances  of  
100s   of   metres)   and   a   lot   of   the   existing   data   occurs   in   areas   for   which   recorded  
ground  motion  is  not  available.    In  this  case,  there  are  two  choices  that  can  be  made:  
1)  the  damage  can  be  related  to  the  observed  macroseismic  intensity  estimates,  2)  
GMPEs  can  be  used  to  estimate  the  ground  motion.  In  the  former  case,  the  problem  
becomes   circular   as   the   same   damage   data   is   used   to   define   the   intensity   measure  
level   and   the   damage   distribution;   an   alternative   could   be   to   use   Internet-­‐based  

42
Chapter  4  

intensity  such  as  Did  You  Feel  it?,  or  ShakeMaps.  When  GMPEs  are  used,  the  choice  of  
the  intensity  measure  (e.g.  PGA,  PGV,  SA)  and  the  choice  of  the  predictive  equation  
will   influence   the   results,   and   the   uncertainty   and   correlation   in   the   ground   shaking  
should   be   included   in   the   estimation   of   the   ground   motion,   though   due   to   its  
complexity   this   is   often   ignored.   Considering   the   large   uncertainty   in   the   conversion  
between   observed   and   instrumental   intensity   measure,   it   is   important   that   the  
intensity  measure  type  selected  for  the  derivation  of  the  fragility  functions  is  one  for  
which  prediction  equations  are  available,  such  that  the  functions  can  be  used  in  the  
forward  modelling  of  damage  and  loss.      

Once   the   building   damage   data   has   been   prepared,   and   the   ground   motion   levels  
have   been   estimated   for   each   geographical   unit   in   the   dataset,   regression   analysis   is  
needed   to   fit   a   curve   to   the   data   (Figure   4.15).   A   number   of   regression   methods   (e.g.  
linear  least  squares,  nonlinear  least  squares)  can  be  applied,  and  the  best  practice  in  
the  case  of  instrumental  intensity  is  to  trial  different  intensity  measure  types  (from  
different   GMPEs)   and   different   regression   methods   to   find   which   combination  
provides  the  best  fit  to  the  data.    

 
Figure  4.15  Empirical  fragility  functions  for  RC  MRF  (Orsini,  1999)  

There  are  thus  a  number  of  drawbacks  to  empirical  fragility  assessment:  

o When   macroseismic   intensity   is   used   for   the   intensity   measure   type,   there   is  
circularity  in  the  method;  

o To  cover  a  wide  range  of  intensity  measure  levels  and  building  typologies,  it  is  
often   necessary   to   combine   damage   surveys   which   might   rely   on   different  
damage  scales  and  different  building  taxonomies;    

43
Basic concepts of seismic risk

o The   data   is   often   clustered   at   the   low   damage/ground   motion   area,   which   can  
influence  the  regression  analysis;  

o Only   building   types   that   have   experienced   damage   can   be   modelled   (and   thus  
these   methods   are   not   suitable   for   evaluating   new   building   typologies,   or  
retrofitted  buildings,  as  discussed  further  in  Chapter  6);  

o The  level  of  ground  shaking  often  needs  to  be  predicted  (thus  the  functions  are  
not  purely  “empirical”  or  “observational”),  and  the  uncertainty  and  correlation  
in  the  latter  accounted  for;  

o To   use   intensity   measure   types   that   correlate   better   with   structural   response  
(such   as   spectral   acceleration   at   the   fundamental   period)   an   estimation   of   the  
period   of   vibration   of   the   buildings   is   needed,   which   introduces   further  
uncertainty.  

Nevertheless,   when   the   data   is   adequately   processed   and   uncertainties   are  


acknowledged  (even  if  not  explicitly  modelled),  then  empirical  fragility  assessment  
has  the  advantage  of  providing  a  realistic  estimate  of  the  expected  damage.  

The   following   publications   are   recent   attempts   to   model   empirical   fragility  


functions:   Rossetto   and   Elnashai   (2003),   Liel   and   Lynch   (2009),   Rota  et   al.   (2008),  
Sarabandi   et   al.   (2004);   Colombi   et   al.   (2008),   Goretti   and   Di   Pasquale   (2004),   Di  
Pasquale  et  al.  (2005),  Karababa  and  Pomonis  (2010),  Amiri  et  al.  (2007).    

4.4.5. Analytical  Fragility  


The   fragility   of   classes   of   buildings   can   be   modelled   through   numerical   models,  
especially   in   the   case   of   reinforced   concrete   buildings   where   extensive   research   and  
development  in  this  field  has  taken  place  over  the  past  30  to  40  years.    

Numerical  models  attempt  to  represent  the  nonlinear  behaviour  of  buildings  under  
different   loading   conditions.   There   are   two   main   camps   of   nonlinear   structural  
analysis:   plastic   hinge   modelling   (“concentrated   inelasticity”)   and   fibre   element  
modelling  (“distributed  inelasticity”).  In  the  former  method,  the  inelastic  behaviour  
is  concentrated  in  springs  at  the  ends  of  the  elements  of  the  building  (the  beams  and  
columns),   which   is   where   the   inelastic   response   is   normally   concentrated,   and   the  
hysteretic  behaviour  of  these  springs  has  to  be  calibrated  before  the  analysis.  In  the  
latter  method,  the  constitutive  behaviour  of  each  material  type  is  explicitly  modelled  

44
Chapter  4  

(through   appropriate   stress-­‐strain   relationships)   and   the   cross-­‐section   of   the  


element  is  divided  into  a  number  of  fibres  and  the  individual  inelastic  behaviour  of  
each  fibre  is  integrated  across  the  section.  The  member  is  divided  into  a  number  of  
elements,   thus   allowing   the   reproduction   of   plastic   hinges   and   the   spread   of  
inelasticity  along  the  member  length  (see  Figure  4.16).    

G aus s node  B
S ec tion  b

G aus s
S ec tion  a
node  A
σ

B
ε

L /2    3 L /2

ε
R C  S ection Unc onfined   C onfined   S teel  F ibres
C oncrete  F ibres C oncrete  F ibres
 
Figure  4.16  Discretisation  in  fibre  modelling  of  a  typical  reinforced  concrete  member  

A  larger  number  of  a  priori  assumptions  are  required  with  plastic  hinge  modelling,  
and   thus   for   fibre   element   modelling   is   recommended   for   less   experienced  
modellers.    

For   fragility   assessment   of   classes   of   buildings   there   are   a   number   of   issues   that  
need  to  be  addressed  when  creating  a  numerical  model:  

o Should   the   building   typology   be   modelled   in   3D   or   simplified   to   a   2D   model?   In  


general  2D  analyses  are  adopted,  though  in  this  case  torsional  behaviour  will  be  
ignored  and  thus  the  typology  should  be  regular  in  plan  (see  Figure  4.17).  

o How   can   the   variability   in   the   geometrical   and   material   properties   amongst   the  
buildings   found   within   a   given   typology   be   modelled?   Once   the   typical   range   of  
values   for   these   properties   is   obtained,   Monte   Carlo   simulation   can   be   used  
such   that   a   number   of   random   buildings   can   be   generated,   each   with   a   given  
beam   length,   beam   depth,   column   length,   column   depth,   wall   thickness,   infill  
panel   thickness,   concrete   compressive   strength,   yield   strength   (and   so   on)  
obtained  from  the  relevant  distribution.  

45
Basic concepts of seismic risk

o How   many   fibres   should   be   employed   across   the   section,   how   many   elements  
should  the  members  be  divided  into?  This  usually  requires  sensitivity  analysis  to  
understand  the  influence  that  these  modelling  choices  have  on  the  response  of  
the  structure.  

CR F
R CM

Torsional Vibration

 
Figure  4.17  Unfavourable  plan  layouts  inducing  torsional  response  in  buildings  

Once  the  geometry,  material  properties,  and  loading  conditions  have  been  assigned  
in  the  numerical  model  a  choice  needs  to  be  made  on  the  type  of  nonlinear  analysis  
to   be   undertaken.   There   are   two   main   types   of   analysis   of   interest   for   fragility  
assessment:  nonlinear  static  and  nonlinear  dynamic.  

4.4.6. Nonlinear  static  analysis  


Nonlinear  static  analysis  involves  the  introduction  of  forces  or  displacements  at  each  
floor   of   the   building   or   frame   to   laterally   displace   the   structure,   following   which   the  
internal  forces  and  deformations  are  calculated  and  balanced.  The  aim  of  a  nonlinear  
static   analysis   is   to   provide   an   envelope   to   the   outputs   of   nonlinear   dynamic  
analysis   under   increasing   levels   of   intensity   (Figure   4.18).   The   output   of   such  
analysis   is   a   pushover   curve,   which   relates   the   shear   force   at   the   base   of   the   building  
(i.e.  the  force  that  is  counteracting  the  lateral  actions  being  applied  to  the  building)  
to  the  displacement  of  a  given  node  of  the  building,  generally  taken  as  the  roof.  

46
Chapter  4  

120

Max Ba se She ar (kN)


100

80

60

40

20

0
0 0.5 1 1.5 2 2.5 3

drift (%)
7000
base shear (kN)

6000

5000

4000

3000 dynamic
DAP
uniform
2000
triangular

1000

0
0% 1% 2% 3%
total drift

Figure  4.18  Maximum  base  shear  versus  top  displacement  obtained  by  running  many  dynamic  analyses  
within  increasing  intensity  (top),  pushover  curves  with  different  lateral  load  assumptions  compared  
with  the  dynamic  output  (bottom).  

Nonlinear  static  analyses  are  less  time  consuming,  and  more  numerically  stable  than  
dynamic   analysis.   However,   as   shown   in   Figure   4.18,   the   results   can   vary   as   a  
function  of  the  assumptions  made  on  the  application  of  the  lateral  actions.  However,  
the  definition  of  global  damage  states  in  terms  of  displacement  limits  is  easier  with  
pushover   curves.   There   are   methods   to   identify   the   yield   displacement   (passing  
through   60-­‐70%   of   the   ultimate   base   shear),   as   well   as   the   ultimate   displacement  
(when   the   base   shear   drops   to   80%   of   the   maximum   value   after   the   peak).   These  
possible  thresholds  are  presented  in  Figure  4.19.  However,  a  check  on  the  response  
of   the   members   at   a   given   global   limit   state   to   damage   should   be   undertaken   (for  
example   to   check   that   the   assumed   ultimate   displacement   really   does   correspond   to  
the  formation  of  plastic  hinges  in  the  majority  of  the  columns  in  a  single  storey,  as  
shown  in  Figure  4.20b).    

47
Basic concepts of seismic risk

150

100

Base shear (kN)


50

Capacity curve
1st limit state
2nd limit state
3rd limit state
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Displacement at the top (m)  
Figure  4.19  Identification  of  global  damage  states  from  a  pushover  curve  

Δp Δp

θp θp

 
Figure  4.20  Collapse  mechanism  (a)  beam-­‐sway,  (b)  column-­‐sway,  (c)  mixed  mechanism  

Simplified   procedures   have   emerged   in   recent   years   to   replace   the   need   to  


undertake  complicated  nonlinear  analyses  of  2D  frames,  which  attempt  to  represent  
pushover   curves   through   a   number   of   simple   formulae   based   on   structural  
mechanics  principles,  and  calibrated  using  nonlinear  analyses  (Calvi,  1999;  Glaister  
and  Pinho,  2003;  Crowley  et  al.  2004;  Bal  et  al.  2010).  These  equations  include  the  
period  of  vibration  as  a  function  of  the  height  (which  can  be  used  to  obtain  the  initial  
stiffness),   and   the   displacements   at   different   performance   levels   as   a   function   of   the  
geometrical  and  material  properties  and  the  mechanism  that  is  assumed  to  form  in  
the  building  (Figure  4.20).  The  collapse  mechanisms  shown  in  Figure  4.20a  and  c  are  
less   brittle   than   that   shown   in   Figure   4.20b   and   occur   at   higher   levels   of   lateral  
deformation.   Older   buildings   or   informal   construction,   that   has   not   been   adequately  
designed   to   withstand   seismic   actions,   is   more   likely   to   fail   with   a   column-­‐sway,  
brittle   mechanism.   A   methodology   for   fragility   derivation   based   on   this   approach  
(Silva  et  al.  2013)  is  presented  in  the  following  section.  

48
Chapter  4  

Once   pushover   curves   have   been   obtained   (either   through   complex   nonlinear  
modelling   or   simplified   equations),   the   response   of   the   building   of   frame   to  
increasing  levels  of  seismic  intensity  needs  to  be  estimated.  There  are  a  number  of  
methods   to   estimate   the   nonlinear   response,   often   termed   Nonlinear   Static  
Procedures   (NSPs),   which   the   main   ones   being   N2   (Fajfar   and   Fischinger,   1988;  
Fajfar   and   Gaspersic,   1996),   Capacity   Spectrum   Method,   with   the   modifications   of  
FEMA  440  (ATC,  2005),  Displacement  Coefficient  Method,  with  the  modifications  of  
FEMA  440  (ATC,  2005).  The  key  steps  of  all  of  these  methods  include:  

1. The  bilinearisation  of  the  pushover  curve;  

2. The  representation  of  the  structure  as  a  single-­‐degree-­‐of-­‐freedom  system  (as  


opposed  to  a  multi-­‐degree-­‐of-­‐freedom  structure);  

3. The   calculation   of   the   nonlinear   response   of   the   SDOF   system   using   an  


elastic  response  spectrum.  

Once   the   response   of   the   structure   to   a   given   accelerogram   (represented   in   terms   of  


a  response  spectrum)  is  found  with  one  of  the  aforementioned  methods,  the  damage  
state  of  the  structure  can  be  found.  Fragility  functions  can  be  generated  as  follows:  

1. Random   generation   of   a   population   of   2D   frames   through   Monte   Carlo  


simulation;  

2. Pushover  curve  for  each  frame,  and  representation  as  the  curve  for  a  SDOF  
system;  

3. Estimate   nonlinear   response   for   each   frame,   using   a   large   selection   of  


records  and  a  given  Nonlinear  Static  Procedure;  

4. Global  damage  state  identification  based  on  the  nonlinear  response;  

5. For  each  intensity  measure  level,  plot  the  cumulative  percentage  of  buildings  
in  each  global  damage  state;  

6. Regression   analysis   to   calculate   the   parameters   (mean   and   standard  


deviation)   of   the   fragility   functions   (can   usually   be   assumed   to   be   a  
lognormal  distribution);  

The   results   of   fragility   functions   obtained   with   nonlinear   static   and   nonlinear  
dynamic  methods  are  likely  to  be  different,  mainly  due  to  the  uncertainty  in  the  NSP,  
which   is   not   accounted   for   in   the   method   described   above.   In   order   to   include   the  

49
Basic concepts of seismic risk

latter   uncertainty   in   the   analysis,   a   number   of   nonlinear   dynamic   analyses   are  


required   in   order   to   estimate   the   uncertainty   in   the   dynamic   response,   given   the  
estimated  static  response  (see  Figure  4.21).  This  uncertainty  (logarithmic  standard  
deviation)   can   then   be   combined   with   the   standard   deviation   calculated   in   point   6  
above.  

 
Figure  4.21  Comparison  of  nonlinear  static  response  estimated  with  the  FEMA  440  method  and  with  
nonlinear  dynamic  analysis  for  7  reinforced  concrete  frames  

For  this  course,  three  nonlinear  static  procedures  will  be  described  in  depth.  These  
methods   vary   considerably   in   terms   of   manner   in   which   the   uncertainties   and  
ground   shaking   are   handles,   and   thus   each   one   of   its   own   limitations   and  
advantages.    

4.4.6.1. SPO2IDA  (Vamvatsikos  and  Cornel,  2006)  


The   Static   PushOver   to   Incremental   Dynamic   Analysis   (SPO2IDA)   is   a   simplified  
methodology,   which   combines   the   results   of   incremental   dynamic   analysis   (IDA)  
with   pushover   analysis.   The   procedure   was   originally   developed   by   Vamvatsikos  
and   Cornell   (2006)   and   has   been   recommended   in   ATC-­‐58   (FEMA   P-­‐58,   2012).  
SPO2IDA  uses   empirical  relationships  from  a  large  database  of  incremental  dynamic  
analysis   results   to   convert   static   pushover   curves   into   probability   distributions   for  
building   collapse   as   function   of   ground   shaking   intensity.   The   process   of   SPO2IDA  
uses  only  one  ground  motion  record  for  a  given  site,  scaled  up  or  down  to  reflect  the  
full   range   of   demands   for   different   return   periods,   of   acceleration   or   displacement  

50
Chapter  4  

amplitudes.   An   Excel   workbook   application   tool   of   SPO2IDA   is   available   for  


download  from  the  ATC-­‐58  project  website1.  

It  should  be  noted  that  this  procedure  is  best  suited  to  low-­‐rise  buildings  of  regular  
configuration.   The   reliability   of   median   values   derived   from   this   procedure   is  
dependent   upon   the   quality   of   analyses   performed   and   the   number   of   intensities  
evaluated.   The   calculation   of   median   capacity   in   terms   of   spectral   acceleration,  
accounting   for   record-­‐to-­‐record   variability,   can   be   conducted   with   the   SPO2IDA   tool  
(Excel  Workbook).  

SPO2IDA   is   applicable   to   any   form   of   capacity   curve,   either   bilinear   or   multilinear  


elasto-­‐plastic.  Given  the  idealized  capacity  curve,  the  SPO2IDA  tool  uses  an  implicit  
R-­‐µ-­‐T   relation   to   correlate   nonlinear   displacement   expressed   in   terms   of   ductility  
(𝜇!"! = ∆!"! /∆! ,   where   ∆! = ∆!"! ),   to   the   corresponding   medians   capacity   in   terms  

of  spectral  acceleration,   Ŝ a ,dsi (T1 ),  as  shown  in  Figure  4.22.  These  medians  capacity  

are   defined   as   the   values   at   which   a   given   damage   threshold   has   been   attained   or  
overcome   by   50%   of   the   analysis   conducted   with   the   ground   motion   pairs.   The  
record-­‐to-­‐record   dispersion   𝛽!   is   estimated   by   considering   the   number   of   records  
that  have  been  used  in  the  SPO2IDA  spreadsheet  (𝛽!  is  estimated  based  on  a  large  
empirical  database  of  incremental  dynamic  analysis  results).  

Figure  4.22  Sample  SPO2IDA  results  of  normalized  spectral  acceleration  at  different  damage  state  
thresholds  

The  steps  required  to  calculate  the  medians  capacity  using  the  SPO2IDA  procedure  
are  described  below:    

1
https://www.atcouncil.org/Projects/atc-58-project.html

51
Basic concepts of seismic risk

1. Develop   an   appropriate   mathematical   model   of   the   building   for   nonlinear   static  


analysis.  Identify  primary  and  secondary  elements  or  components;  define  your  
nonstructural  elements;  foundation  flexibility;  gravity  loads;  and  P-­‐Delta  effects.    

2. Run   a   static   pushover   analysis;   and   make   sure   that   the   analysis   is   extended   to   a  
deformation   such   that   collapse   is   judged   to   occur.   Construct   a   force-­‐
displacement   relationship   considering   different   damage   states   thresholds  
(capacity  curve).  

3. Provide   additional   input   information:   building   weight,   building   height,   and  


fundamental  period.  

4. Fit   the   force-­‐displacement   relationship   into   an   appropriate   idealized  


representation  by  supplying  the  following  reference  control  points:  1-­‐  the  yield  
point;   2-­‐   point   of   peak   strength;   3-­‐   the   onset   of   residual   strength   response;   4-­‐  
the  ultimate  deformation  at  collapse.  

5. Convert   the   fitted   force-­‐displacement   relationship   to   normalized   spectral  


acceleration   and   displacement/ductility   parameters   to   allow   selection   and  
calculation  of  SPO2IDA  input.  

6. Execute   SPO2IDA   and   extract   the   median   capacity   associated   to   each   damage  
threshold.   These   medians   capacity   are   defined   as   the   values   corresponding   to  
the   attainment   or   overcome   of   the   threshold   value   by   50%   of   the   analysis  
conducted  with  ground  motion  pairs  (see  Figure  4.24).  

A  fragility  model  developed  using  the  methodology  was  developed  by  in  Chaulagain  
et  al.  2015  for  reinforced  concrete  structures,  as  depicted  in  Figure  4.23.    
 
1 .0

0 .8
P roba bility  of  exc eedenc e

0 .6
 

0 .4

 m ode ra te
 e x te ns iv e
0 .2
 da m a g e

0 .0
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0 1 .2 1 .4 1 .6 1 .8 2 .0 2 .2 2 .4 2 .6

S pe c tra l  a c c e le ra tion  (g )  T  =  0 .2 9  (s e c )


 
Figure 4.23 Fragility model for reinforced concrete buildings in Nepal (Chaulagain et al. 2015).

4.4.6.2. The  Capacity  Spectrum  Method  (HAZUS  -­‐  FEMA,  2005)  

52
Chapter  4  

The   capacity   spectrum   method   (CSM)   was   initially   proposed   by   Freeman   et   al.  
(1975),   and   it   represents   a   simplified   methodology   for   many   purposes   such   as   the  
evaluation  of  a  large  inventory  of  buildings,  assessment  of  new  or  existing  structures  
or   to   identify   the   correlation   between   damage   states   and   levels   of   ground   motion.  
ATC-­‐40  (1996)  proposes  three  different  procedures  (A,  B  and  C)  for  the  application  
of   the   Capacity   Spectrum   Method.   However,   procedure   B   adopts   some  
simplifications   that   might   not   always   be   valid   and   procedure   C   has   a   very   strong  
graphical   component,   making   it   difficult   for   systematic   applications.   Hence,  
procedure   A,   which   is   characterized   by   its   intuitiveness   and   simplicity,   is   adopted  
herein.   This   procedure   iteratively   compares   the   capacity   and   the   demands   of   a  
structure,  using  a  capacity  curve  (for  the  simplified  SDOF)  and  a  damped  response  
spectrum,   respectively.   The   ground   motion   spectrum   is   computed   for   a   level   of  
equivalent   viscous   damping   that   is   estimated   as   a   function   of   the   displacement   at  
which   the   response   spectrum   crosses   the   capacity   curve,   in   order   to   take   into  
account  the  inelastic  behaviour  of  the  structure.  Iterations  are  needed  until  there  is  
a  match  between  the  equivalent  viscous  damping  of  the  structure  and  the  damping  
applied  to  the  spectrum.  The  final  intersection  of  these  two  curves  approximates  the  
displacement  response  of  the  structure.    

The   initial   proposal   of   this   method   was   heavily   criticized   due   to   its   tendency   to  
underestimate   the   deformation   of   the   structures.   Thus,   in   FEMA-­‐440   (2005),   some  
modifications   were   proposed   mainly   regarding   the   calculation   of   the   equivalent  
viscous  damping.  Another  aspect  worth  further  investigation  in  this  method  is  how  
well  the  bilinear  curves  represent  the  yielding  point  for  different  response  spectra,  
as  this  value  is  crucial  for  the  calculation  of  the  equivalent  viscous  damping.    

ATC-­‐40   defines   the   slope   of   the   first   segment   of   the   bilinear   curve   based   on   the  
initial   stiffness,   which   in   order   to   respect   the   equal   energy   dissipated   rule   (i.e.   the  
area   under   the   capacity   curve   needs   to   be   equal   to   the   area   under   the   bilinear  
curve),  a  yielding  point  located  in  the  elastic  portion  of  the  capacity  curve  might  be  
obtained.   This   misrepresentation   of   the   yielding   occurs   mostly   when   combining   this  
method  with  response  spectra  with  low  levels  of  ground  motion.  An  alternative  for  
this   procedure   has   been   proposed   in   FEMA-­‐273   (1997),   in   which   the   slope   of   the  
first  segment  of  the  bilinear  is  computed  based  on  the  effective  stiffness,  allowing  a  
more  realistic  shape  of  the  bilinear  curves.  This  aspect  is  shown  in  Figure  4.24  and  
Figure   4.25,   where   the   same   capacity   curve   was   used   against   a   weak   (PGA=0.74  

53
Basic concepts of seismic risk

m/s2)   and   a   strong   (PGA=1.42   m/s2)   accelerograms,   respectively,   for   each  


procedure.   0.16
Sd =0.0226 m Sd =0.0393 m
0.14 y y
Say=0.0625 g Say=0.1028 g
0.16

Spectral acceleration (g)


0.16
0.16
1.6 0.16
0.16
1.6 0.12
Sd =0.0226 m Sd =0.0393 m Sd =0.0353 m
0.14 y 2 y y 2
0.14
0.14 Say=0.6252
=0.0625 m/s
g 0.14Sa =0.1028 g
0.14
1.4 0.1 Sa =0.9271
=0.0927 m/s
g
1.4 y y y

Spectral acceleration (g)

(m/s2)
0.12
(m/s2)

Sdy=0.0353
=0.0373 m
(g)

(g)
m Sdy=0.0373 m
Spectral acceleration (g)

Spectral acceleration (g)


0.12
0.12
1.2 0.12
0.12
1.2 0.08
Sd =0.0373 m Sd =0.0373 m
Say=0.0927
y g y
acceleration

acceleration
0.1 Sa =0.0999
Say=0.0999 gg Sa=0.0999
Sa =0.0999 gg

acceleration
0.1 0.1
acceleration

0.1
1 y 0.1
1 0.06 yy
0.08
0.08
0.08
0.8 0.08
0.08
0.8 0.04
0.06 Capacity curve
Spectral

Spectral
0.06
0.06 0.06
0.06
0.6 0.02 Bilinear curve
0.6

Spectral
Spectral

0.04 Response spectrum


0.04
0.04
0.4 Capacity
0.04
0.04
0.4 curve 0
Capacity
Capacity curve
curve 0.02 Bilinear curve 0 Capacity
Capacity 0.05
curve
curve 0.1 0.15
0.02 Bilinear
Bilinear curve
curve Response
0.02 spectrum Bilinear Spectral displacement (m)
curve
0.02
0.2 0.02
0.2 Bilinear curve
0 Response
Response spectrum
spectrum Response
Response spectrum
spectrum
0 0.05 0.1 0.15
0 00 0.05 0.1 0.15 0 00 0.05 0.1 0.15
0 0.05 0.1 Spectral 0.15
displacement 0(m) 0.05 0.1 0.15
0.15
Spectraldisplacement
Spectral displacement(m)
displacement (m)
(m) Spectraldisplacement
Spectral displacement(m)
displacement (m)
(m)  
Figure 4.24 Bilinear curves according to ATC-40 (left) and FEMA-273 (right) for a “weak”
response spectrum.

  0.16

0.14
0.16
0.16 0.16
(g)

0.16
1.6
0.12 0.16
1.6
Sd =0.0373
=0.0373 mm Sdy=0.0226 m Sdy=0.0393 m
y
Spectral acceleration

=0.9991 gm/s2 0.14 y


0.14 Say=0.0999 0.14 Say=0.0625 g Sa =1.0287 m/s2
1.4
0.14
0.1 y 0.14
1.4 y =0.1028 g
y
(g)(g)

0.12
(m/s2)

(m/s2)

Sdy=0.0353 m
(g)
(g)

Sdy=0.0373 m Sdy=0.0373 m
(g)

0.12
0.12
1.2
0.08 0.12
0.12
acceleration

Sd =0.0373 m 1.2 Sd =0.0373 m


y Say=0.0927 g y
acceleration
acceleration

acceleration
Spectral acceleration

=0.0999 gg
Sa=0.0999 0.1 =0.0999 gg
Sa=0.0999
Sa Sa
acceleration

0.1 0.1
acceleration

0.1
1
0.06 yy 0.1
1 yy
0.08
0.08
0.08
0.8
0.04 0.08
0.08
0.8
Spectral

Capacity curve 0.06


Spectral
Spectral

0.06 0.06
Spectral

0.06
0.6
0.02 Bilinear curve 0.06
0.6
Spectral

Spectral

Response spectrum 0.04


0.04
0.04
0.4
0 0.04
0.04
0.4 Capacity curve
0 0.05 Capacity
0.1
Capacity curve
curve 0.15 0.02 Capacity
Bilinear
Capacity curve
curve
curve
0.02
0.02
0.2 Spectral displacementBilinear
Bilinear curve
(m) curve 0.02
0.02 Bilinear
Response
Bilinear curve
spectrum
curve
0.2
Response
Response spectrum
spectrum 0 Response
Response spectrum
spectrum
0 0.05 0.1 0.15
00 0 00 0.05 displacement 0.1 0.15
00 0.05
0.05 0.1
0.1 0.15
0.15 0 Spectral
0.05 0.1(m) 0.15
Spectraldisplacement
Spectral displacement(m)
displacement (m)
(m) Spectral displacement
Spectral displacement (m)(m)
 
Figure 4.25 Bilinear curves according to ATC-40 (left) and FEMA-273 (right) for a “strong”
response spectrum.

A   significant   discrepancy   is   observed   in   the   bilinear   curves   for   the   “weaker”  


response  spectrum.  In  fact,  in  the  case  where  the  ATC-­‐40  guidelines  were  followed,  
an   equivalent   viscous   damping   equal   to   8.94%   was   obtained,   whilst   in   the   second  
case   where   the   FEMA-­‐273   was   used,   a   lower   equivalent   damping   of   5.91%   was  
attained.   With   regards   to   the   second   situation   in   which   a   stronger   spectrum   is  
employed,  both  methods  seem  to  provide  reasonable  results  (equivalent  damping  of  
13.21%   and   12.30%   for   ATC-­‐40   and   FEMA-­‐273,   respectively).   To   summarize,  
according   to   what   is   prescribed   in   ATC-­‐40,   the   Capacity   Spectrum   Method   can   be  
applied  taking  the  following  steps:  

1. Create   a   numerical   model   of   the   structure   and   apply   lateral   forces   to   the  
structure,  in  order  to  assess  the  relation  between  base  shear  and  a  reference  
node  displacement  (i.e.  pushover  curve);  

54
Chapter  4  

2. Convert  the  obtained   curve  for  the  MDOF,  to  the  equivalent  SDOF  (in  terms   of  
spectral  acceleration  (Sa)  versus  spectral  displacement  (Sd));  

3. Construct   a   bilinear   representation   of   the   capacity   curve,   using   one   of   the  


aforementioned  methods;  

4. Calculate  the  response  spectra  (code-­‐based  or  derived  from  a  ground  motion  
record)  for  an  equivalent  5%  elastic  damping;  

5. Intersect   capacity   spectrum   and   reduced   demand   spectrum,   build   new  


bilinear   capacity   spectrum   for   that   point   and   update   effective   viscous  
damping;  

6. If   the   updated   damping   acceptably   matches   the   previously  


determined/assumed   one,   the   Performance   Point   (PP)   of   the   structure   is  
found,  otherwise,  iteration  is  carried  out  until  convergence  is  reached.  

This   performance   point   is   equivalent   to   what   would   be   obtained   by   subjecting   the  


equivalent   single   degree   of   freedom   to   a   nonlinear   time   history   analysis.   This  
response  (i.e.  performance  point)  can  be  used  to  allocate  the  structure  in  a  damage  
state,  based  on  a  pre-­‐established  set  of  displacement  thresholds.    

This   process   can   be   repeated   several   times   considering   other   ground   motion  
records,  as  well  as  structures  (i.e.  building  class).  The  result  will  be  a  distribution  of  
buildings   in   different   damage   states,   according   to   a   set   of   ground   motion   records  
(analytically  derived  damage  probability  matrix).  These  results  can  then  be  used  to  
derive  a  fragility  model,  as  described  in  the  preceding  sections.  

An  alternative  method  for  modelling  the  fragility  of  structures  with  nonlinear  static  
methods   is   to   calculate   the   nonlinear   response   of   the   median   building   for   a   building  
class   on-­‐the-­‐fly   for   a   given   earthquake   scenario.   The   pushover   curve   of   the   SDOF  
system   for   the   median   building   is   thus   needed.   Pre-­‐computed   fragility   functions  
which   relate   the   nonlinear   response   to   the   probability   of   exceedance   of   different  
damage   states   are   then   used   to   obtain   the   damage   distribution;   the   uncertainty   in  
the  aforementioned  fragility  functions  combines  the  uncertainty  in  the  capacity  and  
the  uncertainty  in  the  response.    This  procedure  is  shown  in  Figure  4.26.  However,  it  
is  noted  that  HAZUS  also  includes  a  list  of  fragility  functions  in  terms  of  peak  ground  
acceleration   that   can   be   used   directly   to   calculate   the   damage   distribution   as   a  
function  of  the  ground  shaking  (using  GMPEs).  

55
Basic concepts of seismic risk

 
Figure  4.26  HAZUS  methodology  for  damage  distribution  assessment  (FEMA,  2005)  

4.4.6.1. DBELA  (Silva  et  al.  2013)  


This   section   presents   a   procedure   to   derive   fragility   functions   for   populations   of  
buildings   that   relies   on   the   displacement-­‐based   earthquake   loss   assessment  
(DBELA)  methodology.  This  method  builds  upon  the  urban  assessment  methodology  
proposed   by   Calvi   (1999),   in   which   the   principles   of   structural   mechanics   and  
seismic   response   of   buildings   were   used   to   estimate   the   seismic   vulnerability   of  
classes   of   buildings.   In   this   method,   the   displacement   capacity   and   demand   for   a  
number  of  limit  states  needs  to  be  calculated.  Each  limit  state  marks  the  threshold  
between  the  levels  of  damage  that  a  building  might  withstand,  usually  described  by  
a  reduction  in  strength  or  by  exceedance  of  certain  displacement/drift  levels.  Once  
these   parameters   are   obtained,   the   displacement   capacity   of   the   first   limit   state   is  
compared  with  the  respective  demand.  If  the  demand  exceeds  the  capacity,  the  next  
limit  states  need  to  be  checked  successively,  until  the  demand  no  longer  exceeds  the  
capacity  and  the  building  damage  state  can   be  defined.   If  the   demand   also   exceeds  
the   capacity   of   the   last   limit   state,   the   building   is   assumed   to   have   collapsed.   This  

56
Chapter  4  

procedure  is  schematically  depicted  in  Figure  4.27,  in  which  the  capacities  for  three  
limit  states  are  represented  by  Δi  and  the  associated  demand  by  Sdi.  

Spectral
Displacement
Sd,1 > Δ1 Moderate damage reached
Sd,2 > Δ2 Heavy damage reached Disp. spectrum for moderate damage LS
Sd,3 < Δ3 Collapse not reached
Disp. spectrum for heavy damage LS
Δ3
Disp. spectrum for collapse LS
Sd,3

Sd,2
Δ2
Sd,1 Δ1

Period
T1 T2 T3  
Figure  4.27  -­‐  Comparison  between  limit  state  capacity  and  the  associated  demand  (adapted  from  Bal  et  
al.,  2010).  

In  this  example,  the  demand  exceeds  the  capacity  in  the  first  and  second  limit  state  
but  not  in  the  third  limit  state,  thus  allocating  the  building  to  the  third  damage  state.  

The  demand  in  this  methodology  is  represented  by  a  displacement  spectrum  which  
can   be   described   as   the   expected   displacement   induced   by   an   earthquake   on   a  
single-­‐degree-­‐of-­‐freedom   (SDOF)   oscillator   of   given   period   and   damping.   Therefore,  
the  displacement  capacity  equations  that  are  derived  must  describe  the  capacity  of  a  
SDOF  equivalent  structure,  and  hence  must  give  the  displacement  capacity  at  a  given  
limit  state  (which  could  be  structural  or  non-­‐structural).  

When   considering   structural   limit   states,   the   displacement   at   the   height   of   the  
centre   of   seismic   force   of   the   original   structure   (HCSF)   can   be   estimated   by  
multiplying   the   base   rotation   by   the   height   of   the   equivalent   SDOF   structure   (HSDOF),  
which  is  obtained  by  multiplying  the  total  height  of  the  actual  structure  (HT)  by  an  
effective  height  ratio  (efh)  (see  Figure  4.28).  

57
Basic concepts of seismic risk

 
Figure  4.28  –  Definition  of  effective  height  coefficient  (Glaister  and  Pinho,  2003).  

Pinho   et   al.   (2002)   and   Glaister   and   Pinho   (2003)   proposed   formulae   for   estimating  
the   effective   height   coefficient   for   different   response   mechanisms.   For   what  
concerns   the   beam   sway   mechanism   (or   distributed   plasticity   mechanism,   as   shown  
in  Figure  4.29),  a  ratio  of  0.64  is  proposed  for  structures  with  4  or  less  storeys,  and  
0.44  for  structures  with  20  or  more  storeys.  For  any  structures  that  might  fall  within  
these   limits,   linear   interpolation   should   be   employed.   With   regards   to   the   column-­‐
sway   mechanism   (or   concentrated   plasticity   mechanism,   as   shown   in   Figure   4.29),  
the   deformed   shapes   vary   from   a   linear   profile   (pre-­‐yield)   to   a   non-­‐linear   profile  
(post-­‐yield).   As   described   in   Glaister   and   Pinho   (2003),   a   coefficient   of   0.67   is  
assumed   for   the   pre-­‐yield   response   and   the   following   simplified   formula   can   be  
applied   post-­‐yield   (to   attempt   to   account   for   the   ductility   dependence   of   the  
effective  height  post-­‐yield  coefficient):  

𝜀! !"! !!!
𝑒𝑓! = 0.67 − 0.17 ∙  
𝜀! !"!

Plastic
hinges

 
Figure  4.29  –  Deformed  profiles  for  beam-­‐sway  (left)  and  column-­‐sway  (right)  mechanisms  (adapted  
from  Paulay  and  Priestley,  1992).  

58
Chapter  4  

The  displacement  capacity  at  different  limit  states  (either  at  yield  (Δy)  or  post-­‐yield  
(ΔLSi))  for  bare  frame  structures  can  be  computed  using  simplified  formulae,  which  
are   distinct   if   the   structure   is   expected   to   exhibit   a   beam-­‐   or   column-­‐sway   failure  
mechanism.  These  formulae  can  be  found  in  Bal  et  al.  (2010)  or  Silva  et  al.  (2013).  

In  order  to  estimate  whether  a  given  frame  will  respond  with  a  beam-­‐  or  a  column-­‐
sway   mechanism   it   is   necessary   to   evaluate   the   properties   of   the   storey.   A  
deformation-­‐based  index  has  been  proposed  by  Abo  El  Ezz  (2008)  which  reflects  the  
relation   between   the   stiffness   of   the   beams   and   columns.   This   index   can   be  
computed  using  the  following  formula:  

ℎ!
𝑙!
𝑅=  
ℎ!
𝑙!

where   lc  stands  for  the  column  length.  Abo  El  Ezz  (2008)  proposed  some  limits  for  
this  index  applicable  to  bare  and  fully  infilled  frame  structures,  as  described  in  Table  
4.4.  

Table  4.4  -­‐  Limits  for  the  deformation-­‐based  sway  index.  


Failure  mechanism  
Building  typology  
Beam  sway   Column  Sway  
Bare  frames   R  ≤  1.5   R  >  1.5  
Fully  infilled  frames   R  ≤  1.0   R  >  1.0  
 

The   displacement   demand   is   initially   represented   by   the   5%   damped   spectra.   The  


displacement   is   computed   for   the   period   at   each   limit   state   and   modified   by   a  
correction   factor   (η),   representative   of   the   equivalent   viscous   damping   and   limit  
state   ductility.   In   EC8   (CEN,   2004),   the   following   equation   is   proposed   for   the  
calculation  of  the  correction  factor:  

10
𝜂=  
5 + 𝜉!"

where  ξeq  stands  for  the  equivalent  viscous  damping,  which  can  be  estimated  using  
the  formula  proposed  by  Priestley  et  al.  (2007)  for  RC  structures  with  a  non  negative  
post-­‐yield  ratio:  

where   μLSi   stands   for   the   ductility   at   the   considered   limit   state   (assumed   as   the   ratio  
between   ΔLSi   and   Δy).   More   accurate   approaches   have   recently   been   proposed   to  
estimate   the   correction   factors   (η),   considering   additional   parameters,   such   as   the  

59
Basic concepts of seismic risk

magnitude   or   source-­‐to-­‐site   distance   (Rezaeian   et   al.,   2012).   The   integration   of   such  


methods  in  the  calculation  of  the  displacement  demand  is  in  the  future  development  
plans  of  the  proposed  methodology.  

𝜇!"! !!
𝜉!" = 0.05 + 0.565  
𝜇!"! 𝜋

With   regards   to   the   calculation   of   the   yield   period   (Ty   in   seconds)   for   bare   frame  
structures,   Crowley   and   Pinho   (2004)   and   Crowley   et   al.   (2008)   proposed   a  
relationship   between   the   period   and   the   total   height   of   0.10HT   and   0.07HT   for  
structures   without   and   with   lateral   load   design,   respectively.   For   infilled   frames,   a  
relation   equal   to   0.06HT   has   been   recommended   by   Crowley   and   Pinho   (2006)   for  
structures  without  lateral  load  design.  The  elongated  period  of  vibration  for  any  of  
the  limit  states  (TLSi)  can  be  computed  using  the  following  formula:  

𝜇!"!
𝑇!"! = 𝑇!  
1 + 𝛼 ∙ 𝜇!"! − 𝛼

where   α   stands   for   the   post-­‐yield   stiffness   ratio.   In   cases   where   this   ratio   can   be  
assumed   as   zero,   the   relation   between   TLSi   and   Ty   will   depend   purely   on   the   limit  
state  ductility  as  follows:  

𝑇!"! = 𝑇! 𝜇!"!  

In   this   methodology   to   derive   fragility   functions,   a   randomly   generated   population  


of   buildings   is   produced,   according   to   the   probabilistic   distribution   of   a   set   of  
material  and  geometrical  properties  (e.g.  Bal  et  al.  (2008),  Silva  et  al.  2014b).  Using  
the   set   of   synthetic   buildings,   the   displacement   capacity   Δi   is   calculated   using   the  
equations  described  above.  Similarly,  the  displacement  demand  Sdi  is  also  computed  
for   each   limit   state   period   using   damped   spectra   at   a   level   of   equivalent   viscous  
damping,   from   a   set   of   accelerograms.   However,   if   a   user   wishes   to   avoid   this  
additional   computational   effort,   the   displacement   spectra   can   also   be   provided  
directly.   The   displacement   demand   for   each   limit   state   is   then   computed   by  
modifying   the   elastic   displacement   spectrum   by   a   correction   factor   ηi,  
representative  of  the  equivalent  viscous  damping  and  limit  state  ductility.  

Once   the   capacity   and   demand   displacements   for   the   whole   group   of   synthetic  
buildings   are   computed,   both   sets   of   displacements   are   compared,   and   used   to  
allocate   each   building   in   a   certain   damage   state.   Thus,   for   each   ground   motion  

60
Chapter  4  

record,  percentages  of  buildings  in  each  damage  state  can  be  obtained  and  fragility  
curves   can   be   extrapolated.   In   Figure   4.30,   the   building   damage   distributions   for   4  
records  with  different  levels  of  spectral  acceleration  at  the  fundamental  period  are  
presented.  

       Record  1  –  Sa  0.10  g          Record  2  –  Sa  0.20  g            Record  3  –  Sa  0.30  g          Record  4  –  Sa  0.45  g

 
Figure  4.30  -­‐  Derivation  of  fragility  curves  based  on  building  damage  distribution.  

The   logarithmic   mean,   λ,   and   logarithmic   standard   deviation,   ζ,   that   are   estimated  
for  each  fragility  curve  will  have  some  uncertainty,  due  to  the  scatter  of  the  results.  
Hence,  a  sampling  method  can  be  used  to  properly  evaluate  the  uncertainty  on  the  
statistics.   This   method   can   be   a   continuous   bootstrap   sampling   with   replacement  
from  the  original  dataset  (Wasserman,  2004).  Figure  4.31  shows  an  example  of  the  
estimated  limit  state  exceedance  probability  from  250  records.  

 
Figure 4.31 - Statistical treatment of the parameters defining the curve.

61
Basic concepts of seismic risk

This  methodology  to  derive  fragility  functions  is  summarized  in  Figure  4.32.  

 
Figure  4.32  -­‐  Workflow  of  the  DBELA  fragility  functions  calculator.  

4.4.7. Nonlinear  dynamic  analysis  


Nonlinear  dynamic  analysis  involves  the  introduction  of  accelerograms  at  the  base  of  
the   model   and   the   nonlinear   structural   package   can   be   used   to   compute   the  
response  of  the  building  or  frame  in  terms  of  a  number  of  parameters  such  as:  the  
floor  displacements  or  accelerations  or  the  relative  displacement  between  floors  or  
the   member   forces   and   deformations,   at   any   given   time   in   the   analysis.   Nonlinear  
analyses   can   be   numerically   unstable   and   it   can   be   difficult   to   derive   meaningful  
results   without   significant   modelling   knowledge.   For   fragility   assessment,   a   large  
number   of   accelerograms   with   increasing   intensity   need   to   be   run   through   the  
model   to   capture   the   record-­‐to-­‐record   variability.   For   fragility   assessment,   the  
response   of   the   structure   (in   terms   of   forces   and   deformations   as   mentioned  
previously)   needs   to   be   related   to   the   performance   in   terms   of   a   global   damage  
state.   Experimental   tests   of   members   can   be   useful   for   calibration   of   the  
aforementioned  relationship  between  response  and  damage,  but  the  main  difficulty  
lies   in   defining   how   many   members   should   reach   a   given   level   of   response   for   the  

62
Chapter  4  

definition   of   the   global   damage   state.   Attempts   have   been   made   to   combine   the  
member  responses  into  a  single  global  damage  index  e.g.  Park  and  Ang,  1985).  

12,000

10,000

8,000

6,000

4,000
Acceleration

2,000

-2,000

-4,000

-6,000

-8,000

0 5 10 15
Time

12,000
10,000
8,000
6,000
4,000
2,000
Acceleration

0
-2,000
-4,000
-6,000
-8,000
-10,000
-12,000
-14,000

0 5 10 15
Time

Figure  4.33  Numerical  model  of  a  RC  MRF  for  nonlinear  dynamic  analysis  

The   following   steps   can   thus   be   followed   to   produce   fragility   functions   with  
nonlinear  dynamic  analysis:  

1. Random   generation   of   population   of   2D   frames   through   Monte   Carlo  


simulation;  

2. Nonlinear   dynamic   analysis   for   each   frame   using   a   large   selection   of   ground  
motion  records;    

3. Identification  of  response  and  associated  global  damage  state  for  each  ground  
motion  records;  

4. For  each  intensity  measure  level,  plot  the  cumulative  percentage  of  buildings  
in  each  global  damage  state;  

5. Apply   regression   analysis   to   calculate   the   parameters   (mean   and   standard  


deviation)   of   the   fragility   functions   (usually   assumed   to   follow   a   cumulative  
lognormal  distribution).  

Figure   4.34   illustrates   a   fragility   model   developed   by   Silva   et   al.   (2014b)   for  
moment-­‐frame  reinforced  concrete  structures  using  nonlinear  dynamic  analyses.  

63
Basic concepts of seismic risk

0.9

0.8

Probability of exceedance
0.7

0.6

0.5

0.4

0.3

0.2 Slight
Moderate
0.1 Extensive
Collapse
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Spectral acceleration (g)
 
Figure  4.34  Fragility  function  obtained  from  nonlinear  dynamic  analyses  (Silva  et  al.  2014b)  

The   following   studies   have   used   nonlinear   dynamic   analysis   to   produce   fragility  
functions:   Singhal   and   Kiremidjian   (1996),   Dumova-­‐Jovanoska   (2004)   and   Silva   et  
al.  (2014b).  

There   are   important   drawbacks   in   analytical   fragility   function   assessment   which  


need  to  be  recognized:  

o To   reduce   the   complexity   of   analyses,   buildings   are   often   simplified   as   2D  


regular   models   and   thus   ignore   the   irregularities,   structural   deficiencies   and  
construction  defects  that  cause  the  majority  of  structural  collapses;  

o The   relationship   between   the   modelled   response   and   the   actual   damage   of  
structures   is   not   straightforward   and   often   requires   experimental   testing   for  
improved  calibration.  

Nevertheless,   analytical   fragility   assessment   allows   for   more   control   on   the  


uncertainties   being   modelled,   easier   interpretation   of   the   results,   models   for   new  
and   retrofitted   buildings   can   be   developed,   and   models   for   structures   which   have  
not  experienced  damage  can  be  produced.    

4.5. Consequence  Functions  


As   previously   mentioned,   consequence   functions   (also   known   as   damage-­‐to-­‐loss  
conversion  equations),  allow  fragility  functions  to  be  transformed  into  vulnerability  
functions.  These  functions  are  generally  based  on  observations  (empirical)  or  expert  
opinion.  

In   order   to   transform   fragility   functions   into   vulnerability   functions   in   terms   of  


repair   cost,   so-­‐called   damage   factors   (ratios   of   cost   of   repair   to   cost   of   replacement)  

64
Chapter  4  

are  needed  for  the  considered  building  typology  for  each  damage  state  considered  in  
the   fragility   function.   A   number   of   studies   have   proposed   damage   factors   for  
different   damage   states   (generally   based   on   damage   cost   data   from   previous  
earthquakes),   though   only   a   few   have   looked   at   the   uncertainty   in   these   damage  
factors   for   a   given   damage   state.   Masi   et   al.   (2002)   represents   this   variation   as   a  
Beta   distribution,   based   on   post-­‐earthquake   survey   data   from   Italy,   whilst   Goretti  
and  Di  Pasquale  (2004)  assume  normal  distributions  for  the  damage  factors.    

Construction   practices   and   costs,   in   particular   those   applied   in   the   repair   and  
reconstruction   phase,   vary   significantly   from   one   country   to   another,   and   thus  
damage   factors   should   vary   accordingly.   Assigning   repair   types   and   associated   costs  
to   damage   states   depends   on   a   number   of   factors   from   failure   mechanism,   to   size  
and   geometry   of   buildings   in   the   building   class,   to   engineering   practice   of   the  
country.   Some   examples   of   attempts   to   calculate   damage   factors   include   Bal   et   al.  
(2008)  and  Hill  and  Rossetto  (2008b,  c).    

For   the   method   proposed   by   Giovinazzi   and   Lagomarsino   (2004,   2006),   a  


correlation   between   the   mean   damage   grade   μD   and   what   is   referred   to   as   the  
damage  index  DI  (i.e.  the  ratio  between  the  repair  and  the  reconstruction  cost)  has  
been  proposed,  which  has  been  fit  to  data  present  in  the  literature  (though  it  is  not  
clear  whether  this  is  purely  empirical  or  a  mixture  of  expert  opinion  and  observed  
data):  

µD  =  5  DI0.57  

Hence,   the   curves   presented   in   Figure   4.13   can   be   converted   from   mean   damage  
grade   to   loss   ratio   (ratio   of   cost   of   repair   to   cost   of   replacement),   and   thus  
vulnerability  functions  can  be  produced.    

The   definition   of   replacement   cost   varies   in   the   aforementioned   studies,   from   “new”  
build   cost   (Bal   et   al.   2008)   to   market   value   of   the   property   before   the   earthquake  
(Whitman   et   al.   1973),   to   construction   replacement   cost   (FEMA,   2003).   It   is   noted  
that   whatever   is   the   definition,   it   should   be   compatible   with   the   definition   of  
replacement  cost  in  the  exposure  model.  

There   are   thus   a   number   of   issues   that   should   be   considered   when   using   existing  
damage  factors:  

o Are  the  damage  states  equivalent  to  those  of  the  fragility  functions?  

65
Basic concepts of seismic risk

o What  is  the  meaning  of  the  replacement  cost  used  in  their  derivation?  

o Are  they  for  the  same  country,  or  is  the  engineering  and  construction  practice  
similar?  

Coburn   and   Spence   (2002)   discuss   the   many   earthquake-­‐induced   causes   of   death,  
which   include   fires   following   earthquakes,   tsunamis,   rockfalls   and   landslides.  
However,  it  is  clear  that  in  most  large-­‐scale  earthquake  disasters  the  main  cause  of  
death  is  the  collapse  of  buildings    (see  Figure  4.35).  Over  the  last  century  about  75%  
of  fatalities  attributed  to  earthquakes  have  been  caused  by  the  collapse  of  buildings  
(Coburn   and   Spence,   2002),   and   the   greatest   proportion   of   victims   die   in   the  
collapse   of   masonry   buildings.   Unfortunately   these   buildings   are   very   common   in  
seismic   areas   and   make   up   a   large   proportion   of   the   building   stock   around   the  
world.   It   should   be   noted,   however,   that   in   some   areas   of   the   world   the   buildings  
constructed  in  reinforced  concrete,  which  has  become  the  material  of  choice  for  new  
construction,  are  actually  highly  vulnerable  and  when  they  do  collapse  can  be  more  
lethal   and  cause   the   death   of   a   higher   percentage   of   their   occupants   than   masonry  
buildings.  Figure  4.35  shows  how  there  has  been  an  increase  in  fatalities  caused  by  
the  collapse  of  reinforced  concrete  buildings  in  the  second  half  of  the  last  century.  

 
Figure  4.35  Breakdown  of  earthquake-­‐induced  fatalities  by  cause  (Coburn  and  Spence,  2002)  

Coburn  and  Spence  (2002)  provide  a  plot  of  the  total  number  of  people  killed  versus  
the  total  number  of  heavily  damaged  buildings  (Figure  4.36),  where  it  can  be  readily  
seen   that   deaths   can   be   correlated   to   the   destruction   caused   by   earthquakes.  
However,  it  should  be  noted  that  the  casualty  values  are  highly  dispersed  when  less  
than  5000  buildings  were  damaged.  

66
Chapter  4  

 
Figure  4.36  Relationship  between  the  number  of  fatalities  and  the  number  of  buildings  damaged  
heavily  in  earthquake  (Coburn  and  Spence,  2002)  

In   order   to   transform   fragility   functions   into   vulnerability   functions   in   terms   of  


casualties  and  fatalities,  casualty  loss  ratios  are  needed  which  relate  the  number  of  
casualties  to  the  number  of  occupants  of  buildings  of  a  given  typology  damage  to  a  
given   damage   state.   Detailed   post-­‐earthquake   consequences   data   is   needed   in   order  
to  anchor  casualty  loss  ratios  to  the  damage  and  building  type,  and  this  information  
is   often   not   available.   Furthermore,   it   is   difficult   to   extract   the   death   and   injury  
counts   that   are   due   to   ground   shaking   alone.   Despite   these   difficulties,   Spence  
(2007)   and   So   (2009)   have   attempted   to   produce   casualty   loss   ratios   as   described  
previously   (Figure   4.4)   and   recent   earthquakes   have   proved   to   have   extremely  
detailed   casualty   information   (e.g.   2011   New   Zealand   earthquake:   So,   personal  
communication).  

Table  4.5  Injury  distributions  for  specific  building  types  (Spence,  2007).  UI=uninjured;  I1=slight  
injuries;  I2=moderate  injuries;  I3=serious  injuries;  I4=critical  injuries;  I5=deaths  
  Complete  Damage  State  
UI   I1   I2   I3   I4   I5  
Masonry  (1F)   23.6%   50.0%   12.0%   8.0%   0.4%   6.0%  
Masonry  (2-­‐3F)   16.5%   50.0%   15.0%   10.0%   0.5%   8.0%  
Masonry  (≥4F)   9.4%   50.0%   18.0%   12.0%   0.6%   10.0%  
RC  (1F)   32.9%   30.0%   19.0%   3.0%   0.2%   15.0%  
RC  (2-­‐3F)   20.8%   30.0%   23.0%   4.0%   0.2%   22.0%  
RC  (≥4F)   9.7%   30.0%   27.0%   5.0%   0.3%   28.0%  
 

67
Chapter  5  
Basic  concepts  of  seismic  risk    
This   Chapter   describes   the   convolution   of   seismic   hazard,   exposure   and  
vulnerability  for  the  calculation  of  physical  risk.  The  following  sections  focus  on  the  
methods  that  have  been  implemented  within  the  open-­‐source  software  OpenQuake  
(Silva  et  al.  2014c)  that  is  being  developed  by  the  Global  Earthquake  Model.    

5.1. Scenario  Risk  Assessment  


Scenario  risk  assessment  involves  the  calculation  of  losses  and  loss  statistics  due  to  
a  single,  deterministic  earthquake  (such  as  a  large  damaging  historical  earthquake),  
for   a   collection   of   assets.   Such   analyses   are   of   importance,   for   example,   for  
understanding   the   validity   of   a   given   model,   for   emergency   management   planning,  
for  raising  societal  awareness  of  risk,  or  for  rapid  post-­‐earthquake  loss  assessment.  

This   workflow   requires   the   definition   of   a   finite   rupture   (rupture   scenario   model),  
an   exposure   model   and   a   physical   vulnerability   model.   The   characteristics   of   the  
rupture  are  used  to  generate  ground  motion  fields  (i.e.  estimates  of  the  intensity  of  
ground  shaking  at  each  location  in  the  model),  considering  the  aleatory  variability  in  
the  GMPE  or  IPE  with  or  without  spatial  correlation.  For  a  given  ground  motion  field,  
the  intensity  measure  level  at  a  given  site  is  used  to  obtain  a  random  sample  of  the  
loss  ratio  (considering  the  mean  and  standard  deviation  of  the  loss  ratio  at  the  given  
intensity  measure  level  from  the  vulnerability  model)  for  each  asset  contained  in  the  
exposure  model.  Using  the  sampled  loss  ratios  for  each  asset,  the  mean  and  standard  
deviation   of   the   loss   ratios   across   all   ground   motion   fields   can   be   calculated.   Loss  
ratios  are  converted  into  losses  by  multiplying  by  the  value  of  the  asset  given  in  the  
Chapter  5  

exposure  model.  It  is  furthermore  possible  to  sum  the  losses  throughout  the  region  
and   to   compute   the   mean   and   standard   deviation   of   the   total   loss,   and   to   plot  
statistics   such   as   the   mean   loss   on   maps,   as   presented   in   Figure   5.1   for   the   Pisco  
(Peru)  earthquake  of  2007.  

 
Figure  5.1  Mean  economic  loss  map  for  the  Pisco  (Peru)  earthquake  of  August  2007.  

5.2. Scenario  Damage  Assessment  


Scenario  damage  assessment  leads  to  the  calculation  of  the  distribution  of  buildings  
in   various   damage   states   (e.g.   no   damage,   moderate,   collapse),   given   a   single  
earthquake.   This   module   requires   the   definition   of   a   finite   rupture,   an   exposure  
model   and   a   fragility   model.   The   main   results   of   general   interest   are   the   damage  
distribution  per  asset,  total  damage  distribution  and  collapse  maps.  This  calculator  
is   very   similar   to   the   Scenario   Risk   calculator,   however   instead   of   a   vulnerability  
model,   a   fragility   model   is   required.   Once   the   ground   motion   field   have   been  
generated,  the  damage  distribution  for  each  asset  in  each  location  is  obtained  from  
the   appropriate   fragility   model   (as   a   function   of   the   building   typology).   Using   the  
number   or   area   of   each   building   typology   in   each   location,   the   damage   distributions  
can   be   weighted   and   an   aggregated   damage   distribution   for   the   whole   region   is  
obtained.   The   mean   and   standard   deviation   of   the   aggregated   damage   distribution  
across   the   ground   motion   fields   can   also   be   obtained.   This   type   of   results   is  
illustrated  in  Figure  5.2.  

69
Basic concepts of seismic risk

Number#of#buildings#

Collapse#
Extensive#
Moderate#
Slight#
Adobe#
Wooden# No#damage#
Masonry#
Stone#
RC#

                         
Figure  5.2  Distribution  of  damage  per  building  typology  (left)  and  mean  collapse  map  (right)  for  the  
Pisco  (Peru)  earthquake  of  August  2007.  

5.3. PSHA  and  Event-­‐based  Risk  Assessment  


The   aforementioned   types   of   risk   analysis   use   single,   pre-­‐defined   events   and   are  
thus   conditional   on   that   event   occurring.   In   order   to   consider   the   risk   considering  
any  possible  event  that  could  occur  within  a  given  region,  the  probability  of  future  
events  needs  to  be  incorporated  into  the  analysis.  There  are  two  ways  in  which  this  
can  be  done:  

o By   using   the   output   of   Probabilistic   Seismic   Hazard   Assessment   (PSHA)   in  


terms   of   hazard   maps   and   curves   (known   as   classical   PSHA-­‐based   risk  
assessment  herein).  

o By   using   stochastic   event   sets,   generated   from   seismic   hazard   input   models  
(known  as  probabilistic  event-­‐based  risk  assessment  herein).  

When  the  risk  assessment  only  needs  to  provide  mean  outputs  (e.g.  average  annual  
losses,   AAL)   or   relative   estimates   of   risk,   then   the   former   approach   is   appropriate.  
This   method   is   commonly   used   for   ranking   countries   in   terms   of   risk   or   for  
producing  risk  maps  of  AAL.  When  the  full  range  of  probabilities  of  losses  need  to  be  
robustly  estimated  (through  loss  exceedance  curves),  then  the  latter  method  is  more  
appropriate.  

As  with  the  scenario-­‐based  risk  calculator,  the  main  input  models  are  exposure  and  
vulnerability.  Hazard  curves  are  the  output  of  the  hazard  module  that  are  then  input  

70
Chapter  5  

into   the   risk   calculator.   These   hazard   curves   can   be   combined   with   a   vulnerability  
and   exposure   model   to   derive   asset-­‐specific   loss   exceedance   curves.   Due   to   the  
employment   of   logic   trees,   it   is   possible   to   have   multiple   hazard   curves   at   each  
location   (one   hazard   curve   per   logic   tree   branch).   Thus,   various   loss   exceedance  
curves  are  also  calculated  for  each  asset,  and  a  user  has  the  possibility  of  extracting  
results   from   a   specific   logic   tree   branch,   the   mean   loss   exceedance   curves,   or   loss  
curves  corresponding  to  a  specific  percentile.  These  loss  exceedance  curves  can  be  
used  to  derive  risk  maps,  containing  the  expected  loss  for  a  given  return  period,  as  
presented  in  Figure  5.3.  

Figure  5.3  Economic  losses  for  a  probability  of  exceedance  of  10%  (left)  and  2%  (right)  in  50  years.  

The  same  exposure  and  vulnerability  models  mentioned  previously  are  also  valid  for  
the   probabilistic   event-­‐based   risk   assessment   approach.   As   with   the   scenario   risk  
calculator,  ground  motion  fields  (GMFs)  are  used  in  the  risk  calculations,  but  a  much  
larger   number   of   GMFs   are   used;   one   per   event   in   the   stochastic   event   set.     For   each  
ground   motion   field,   the   intensity   measure   level   at   a   given   site   is   combined   with   a  
vulnerability   function,   from   which   a   loss   ratio   is   randomly   sampled   for   each   asset.  
Spatial   correlation   of   the   ground   motion   residuals   and   vulnerability   correlation   in  
the  loss  ratios  within  assets  of  the  same  taxonomy  can  be  accounted,  as  described  in  
the   Scenario   Risk   Calculator.   The   distribution   of   loss   per   asset   across   all   ground  
motion  fields  is  used  to  estimate  the  rate  of  exceeding  a  number  of  loss  thresholds,  
which   can   then   be   converted   into   a   loss   exceedance   curve   assuming   a   Poissionian  
distribution.  Once  again,  a  logic  tree  structure  can  be  employed  in  the  calculation  of  

71
Basic concepts of seismic risk

the   ground   motion   fields,   which   leads   to   a   number   of   loss   exceedance   curves   per  
asset  equal  to  the  number  of  branches  considered  in  the  hazard  calculations.  These  
loss   curves   can   be   for   a   single   asset,   or   aggregated   for   all   assets   in   the   exposure  
model.  Once  again,  loss  maps  can  also  be  generated,  as  illustrated  in  in  Figure  5.3.  

5.4. Retrofitting  Decision  Support  Tool  


Seismic   retrofitting   involves   structural   interventions   that   aim   to   improve   the  
performance  of  buildings  in  earthquakes.  Conventional  intervention  techniques  for  
retrofitting   can   be   subdivided   into   global   and   member   intervention   types.   Examples  
of   the   former   include   addition   of   RC   structural   walls,   steel   bracing   and   buttresses  
whilst  the  latter  can  take  the  former  of  member  jacketing  or  injection  of  epoxy  resin.  
Figure  5.4  shows  four  different  global  techniques  that  could  be  applied  to  retrofit  a  
building  with  an  open  ground  floor  storey.  Each  method  has  its  pros  and  cons,  which  
need  to  be  considered:  

o What  impact  will  the  scheme  have  on  the  foundations?    

o How  much  impact  will  it  have  on  the  functioning  of  the  building?    

o How  much  does  it  cost?  

o How  much  is  the  performance  of  the  building  improved?  

 
Figure  5.4  Four  global  retrofitting  schemes  for  a  building  with  a  soft  storey.  

72
Chapter  5  

Benefit-­‐cost   analysis   (BCA)   is   a   systemic   procedure   for   evaluating   decisions   that  


have  an  impact  on  society  and  that  can  be  applied  in  the  case  of  retrofitting  to  find  
the  most  attractive  solution  that  maximises  the  net  present  value.  

The  direct  costs  of  the  various  options  (such  as  those  shown  in  Figure  5.4)  need  to  
be   estimated,   considering   also   the   impact   that   the   time   that   the   building   will   be  
unavailable   might   have   on   the   cost.   The   different   retrofitting   schemes   that   can   be  
applied   need   to   be   modelled.   This   is   where   analytical   fragility   assessment   is  
fundamental  in  risk  assessment,  such  that  a  robust  assessment  of  the  impact  of  the  
different  retrofitting  schemes  on  the  response  of  the  building  can  be  estimated.  The  
previously  described  methods  can  be  applied  for  models  with  different  retrofitting  
schemes  and  vulnerability  functions  of  the  original  and  retrofitted  building  derived  
(Figure  5.5).  

 
Figure  5.5  Discrete  vulnerability  functions  (uncertainty  not  plotted)  for  original  and  retrofitted  
buildings  

The  next  step  involves  assessing  the  benefits  of  the  different  retrofitting  schemes  in  
terms  of  reduced  losses.  The  losses  should  be  economical  in  order  to  be  compared  
with   the   cost   of   retrofitting,   but   this   does   not   mean   that   only   repair   cost   can   be  
accounted  for.  Although  a  difficult  and  somewhat  controversial  subject,  a  value  of  life  
can   be   assigned   to   the   number   of   fatalities   as   well.   Economists   have   used   several  
estimation  techniques  for  estimating  the  value  of  life.  These  range  from  hypothetical  
surveys  where  people  are  asked  how  much  they  must  be  paid  to  accept  certain  risks,  
to   examining   the   wage   premium   people   working   in   hazardous   jobs   are   given   to  
compensate  them  for  the  additional  risks  they  are  incurring.  

73
Basic concepts of seismic risk

Loss   exceedance   curves   need   to   be   produced   using   the   classical   PSHA-­‐based   risk  
workflow,   considering   the   original   vulnerability   functions,   and   then   again   with   the  
retrofitted  vulnerability  model.    

 
Figure  5.6  Loss  exceedance  curves  for  original  and  retrofitted  buildings  

The  Benefit-­‐Cost  Ratio  (BCR)  can  then  be  calculated  as  follows:  

(𝐴𝐴𝐿! − 𝐴𝐴𝐿! ) (1 − 𝑒 !!" )


𝐵𝐶𝑅 = .  
𝐶 𝑟

where  AALC  is  the  average  annual  loss  of  the  current  (original)  building,  AALR  is  the  
average   annual   loss   of   the   retrofitted   building,   r   is   the   discount   rate   (to   bring   the  
losses  to  the  future  present  monetary  value),  t  is  the  design  life  of  the  building  (e.g.  
50  years)  and  C  is  the  cost  of  retrofitting  (that  needs  to  be  included  in  the  exposure  
model).  The  difference  in  average  annual  loss  with  and  without  retrofitting  provides  
the   average   avoided   loss   per   year   should   retrofitting   be   employed.   This   is   then  
multiplied  by  a  factor  that  brings  the  avoided  losses  over  the  life  of  the  building  to  
the  present  value.  For  a  given  design  life,  as  the  discount/interest  rate  increases,  the  
future   benefits   decrease   as   the   losses   in   the   future   are   higher,   hence   the   average  
saving   in   loss   over   t   years   (the   benefit)   is   lower   and   the   BCR   decreases.   Discount  
rates  are  generally  assumed  to  be  between  3  and  6%  (FEMA,  1992).  

The   decision   on   whether   to   retrofit   and   which   retrofitting   scheme   to   adopt   can   be  
based  on  the  scheme  with  the  highest  BCR  (above  1).    

74
REFERENCES  
Abo   El   Ezz,   A.   (2008).   “Deformation   and   strength   based   assessment   of   seismic   failure  
mechanisms  for  existing  RC  frame  buildings”.  MSc  Thesis,  ROSE  School,  Pavia,  Italy.  
 
Abrahamson,   N.   A.   (2006).   “Seismic   hazard:   Problems   with   current   practice   and   future  
developments”.   Proceedings   of   the   1st   European   Conference   on   Earthquake   Engineering  
and  Seismology,  Geneva,  Switzerland.  
 
Amiri,  G.G.,  Jalalian,  M.,  Amrei,  S.A.R.  (2007)  “Derivation  of  vulnerability  functions  based  on  
observational   data   for   Iran,”   Proceedings   of   International   Symposium   on   Innovation  
and  Sustainability  of  Structures  in  Civil  Engineering,  Tongji  University,  China.    
 
Applied   Technology   Council   (ATC)   (1985)   ATC-­‐13,   Earthquake   Damage   Evaluation   Data   for  
California,  Applied  Technology  Council,  Redwood  City,  CA,  492  pp.  
 
Applied   Technology   Council   (ATC)   (1996).   ATC-­‐40   Seismic   Evaluation   and   Retrofit   of  
Concrete  Buildings,  Volumes  1  and  2,  Report  No.  ATC-­‐40,  Applied  Technology  
Council,  Redwood  City,  California,  USA.  
 
Bal,   I.,   Crowley,   H.,   Pinho,   R.,   Gulay,   F.   (2008b).   “Detailed   assessment   of   structural  
characteristics  of  Turkish  RC  building  stock  for  loss  assessment  models”,  Soil  
Dynamics  and  Earthquake  Engineering,  28:914-­‐932.  
 
Bal   I.,   Crowley,   H.   and   Pinho,   R.   (2010)   Displacement-­‐Based   Earthquake   Loss   Assessment:  
Method   Development   and   Application   to   Turkish   Building   Stock,   ROSE   Research   Report  
2010/02,  IUSS  Press,  Italy.  
 
Bird,   J.F.   and   Bommer,   J.J.   (2004)   “Earthquake   losses   due   to   ground   failure,”   Engineering  
Geology,  75(2),  pp.  147–79.  
 
Bolt,  B.A.  (1999)  Earthquakes,  Fourth  edition,  W.H.  Freeman.    
 
Bommer,   J.   J.,   Spence,   R.,   Erdik,   M.,   Tabuchi,   S.,   Aydinoglu,   N.,   Booth,   E.,   Re,   D.   D.,  
Pterken,   D.   (2002).   “Development   of   an   Earthquake   Loss   Model   for   Turkish  
Catastrophe  Insurance”.  Journal  of  Seismology,  6:431-­‐446.  
 
Bommer,   J.J.   (2004)   Engineering   Seismology,   Class   notes   for   the   ROSE   School,   IUSS,   Pavia,  
Italy  
 
Bommer,   J.   J.,   Abrahamson,   N.   A.   (2006).   "Why   do   Modern   Probabilistic   Seismic  
Hazard   Analyses   Often   Lead   to   Increased   Hazard   Estimates?".   Bulletin   of   the  
Seismological  Society  of  America,  96:1967-­‐1977.  
 
Bommer,  J.J.  and  Stafford,  P.J.  (2008).  Seismic  hazard  and  earthquake  actions,  In:  Elghazouli,  
A.Y.,  Seismic  Design  of  Buildings  to  Eurocode  8,  Taylor  &  Francis  
 
Brzev   S.,   C.   Scawthorn,   A.W.   Charleson,   L.   Allen,   M.   Greene,   K.   Jaiswal,   and   V.   Silva   (2013),  
GEM  Building  Taxonomy  Version  2.0,  GEM  Technical  Report  2013-­‐02  V1.0.0,  188  pp.,  
GEM  Foundation,  Pavia,  Italy,  doi:10.13117/GEM.EXP-­‐MOD.TR2013.02.  
 
Calvi,  G.M.  (1999)  “A  displacement-­‐based  approach  for  vulnerability  evaluation  of  classes  of  
buildings,”  Journal  of  Earthquake  Engineering,  Vol.  3,  No.  3,  pp.  411-­‐438.  
 
Calvi,  M.,  Pinho,  R.,  Magenes,  G.,  Bommer,  J.,  Restrepo-­‐Vélez,  L.,  Crowley,  H.  (2006).  
“Development   of   Seismic   Vulnerability   Assessment   Methodologies   over   the  
past  30  years“,  ISET  Journal  of  Earthquake  Technology,  43(3):75-­‐104.  
 
Casarotti,   C.,   Pinho,   R.   (2007).   “An   adaptive   capacity   spectrum   method   for  
assessment  of  bridges  subjected  to  earthquake  action”.  Bulletin  of  Earthquake  
Engineering;  5(3):377–390.  
 
CEN   (2004).   Eurocode   8:   Design   of   Structures   for   Earthquake   Resistance   -­‐   Part   1:  
General  rules,  seismic  actions  and  rules  for  buildings.  European  Committee  for  
Standardization,  Brussels,  Belgium.  
 
Chaulagain,   H.,   Rodrigues,   H.,   Silva,   V.,   Spacone,   E.,   Varum,   H.   (2014)   "Earthquake  
Loss   Estimation   for   the   Kathmandu   Valley",   Bulletin   of   Earthquake  
Engineering.  
 
Chopra,  A.  K.,  Goel,  R.  K.,  (2000).  “Evaluation  of  NSP  to  estimate  seismic  deformation:  
SDF  systems”.  Journal  of  Structural  Engineering,  126(4):482-­‐490.  
 
Chopra,   A.   K.,   Goel,   R.   K.   (2002).   “A   modal   pushover   analysis   procedure   for  
estimating   seismic   demands   for   buildings”.   Earthquake   Engineering   &  
Structural  Dynamics;  31(3):561–582.  
 
Coburn,  A.  W.  and  Spence,  R.  (2002)  Earthquake  Protection,  John  Wiley  &  Sons.  
 
Colombi,   M.,   Borzi,   B.,   Crowley,   H.,   Onida,   M.,   Meroni,   F.,   Pinho,   R.   (2008)   “Deriving  
vulnerability   curves   using   Italian   earthquake   damage   data,”   Bulletin   of   Earthquake  
Engineering,  6,  pp.  485-­‐504    
 
Crowley,   H.   and   Pinho,   R.   (2004)   “Period-­‐height   relationship   for   existing   European  
reinforced  concrete  buildings,”   Journal  of  Earthquake  Engineering,  Vol.  8,  Special  Issue  
1.,  pp.93-­‐119  
 
Crowley   H.,   Pinho   R.   and   Bommer   J.J.   (2004)   “A   probabilistic   Displacement-­‐based  
Vulnerability   Assessment   Procedure   for   Earthquake   Loss   Estimation,”   Bulletin   of  
Earthquake  Engineering,  Vol.  2,  pp.  173-­‐219.  
 
Crowley,   H.,   Bommer,   J.J.   and   Stafford,   P.J.   (2008)   “Recent   Developments   in   the   Treatment   of  
Ground-­‐Motion   Variability   in   Earthquake   Loss   Models,   “   Journal   of   Earthquake  

76
 

Engineering,  Vol.  12  (S2),  pp.  71-­‐80.  


 
Crowley,   H.,   Pinho,   R.   (2006).   “Simplified   equations   for   estimating   the   period   of  
vibration  of  existing  buildings”.  Proceedings  of  the  1st  European  Conference  on  
Earthquake  Engineering  and  Seismology,  Geneva,  Switzerland.  
 
Crowley,  H.,  Miriam,  C.,  Borzi,  B.,  Faravelli,  M.,  Onida,  M.,  Lopez,  M.,  Polli,  D.,  Meroni,  
F.  (2009).  "A  comparison  of  seismic  risk  maps  for  Italy".  Bulletin  of  Earthquake  
Engineering,  7(1):149-­‐180.  
 
Cua   et   al.   (2010)   Available   from   URL:  
http://www.globalquakemodel.org/system/files/doc/GEM-­‐TechnicalReport_2010-­‐
4.pdf  
 
Cummins,   J.   D.,   Mahul,   O.   (2009).   Catastrophe   Risk   Financing   in   Developing  
Countries:   Principles   for   Public   Intervention.   The   World   Bank,   Washington  
D.C.,  USA.      
 
Daniell,  J.  E.,  Wenzel,  F.,  Khazai,  B.  (2010).  "The  Cost  of  Historic  Earthquakes  Today  –  
Economic   Analysis   since   1900   through   the   use   of   CATDAT".   Proceedings   of   the  
Australian  Earthquake  Engineering  Society  Conference,  Perth,  Australia.  
 
Dengler,   L.A.   and   Dewey,   J.W.   (1998)   “An   Intensity   Survey   of   Households   Affected   by   the  
Northridge,   California,   Earthquake   of   17   January,   1994,”   Bulletin   of   the   Seismological  
Society  of  America,  Vol.  88,  p.  441-­‐462.  
 
Di   Pasquale,   G.,   Orsini,   G.   and   Romeo,   R.W.   (2005)   “New   developments   in   seismic   risk  
assessment  in  Italy”,  Bulletin  of  Earthquake  Engineering,  Vol.  3,  No.  1,  pp.  101-­‐128.  
 
Douglas,   J.   (2003)   “Earthquake   ground   motion   estimation   using   strong-­‐motion   records:   a  
review   of   equations   for   the   estimation   of   peak   ground acceleration   and   response  
spectral  ordinates,”  Earth-­‐Science  Reviews,  61,  pp.  43-­‐104.  
 
Dumova-­‐Jovanoska,   E.   (2004)   “Fragility   curves   for   RC   structures   in   Skopje   region,”  
Proceedings  of  13th  World  Conference  on  Earthquake  Engineering,  Vancouver,  Canada.  
 
Erdik,  M.,  Aydinoglu,  N.,  Fahjan,  Y.,  Sesetyan,  K.,  Demircioglu,  M.,  Siyahi,  B.,  Durukal,  
E.,   Ozbey,   C.,   Biro,   Y.,   Akman,   H.,   Yuzugullu   O.   (2003).   “Earthquake   Risk  
Assessment   for   Istanbul   Metropolitan   Area”,   Earthquake   Engineering   and  
Engineering  Vibration,  2(1).  
 
FEMA  (1992)  A  benefit/cost  model  for  the  seismic  rehabilitation  of  buildings  (FEMA  227),  Vols  
1,  2.  VSP  Associates,  Sacramento,  California.  
 
FEMA-­‐273   (1997).   NEHRP   Guidelines   for   the   Seismic   Rehabilitation   of   Buildings.  
Report   No.   FEMA   273.   Federal   Emergency   Management   Agency,   Washington  
D.C.,  USA.  
 

77
FEMA   (2003)   HAZUS-­‐MH   MR4   Technical   Manual,   Federal   Emergency   Management   Agency.  
http://www.fema.gov/plan/prevent/hazus/hz_manuals.shtm    
 
FEMA-­‐440   (2005).   Impovement   of   Nonlinear   Static   Seismic   Analysis   Procedures.  
Federal  Emergency  Management  Agency,  Washington  D.C.,  USA.  
 
Gamba,   P.,   Cavalca,   D.   Jaiswal,   K.,   Huyck,   C.,   Crowley,   H.   (2012).   "The   GED4GEM  
Project:   Development   of   a   Global   Exposure   Database   for   the   Global  
Earthquake  Model  Initiative".  Proceedings  of  the  15th  WCEE,  World  Conference  
on  Earthquake  Engineering,  Lisbon,  Portugal.  
 
Freeman,   S.,   Nicoletti,   J.,   Tyrell,   J.   (1975).   “Evaluation   of   Existing   Buildings   for  
seismic   risk   -­‐   A   case   study   of   Puget   Sound   Naval   Shipyard,   Bremerton,  
Washington”.   Proceedings   of   the   1st   U.S.   National   Conference   on   Earthquake  
Engineering,  Berkley,  USA.  
 
Freeman,   S.,   (2004).   “Review   of   the   development   of   the   capacity   spectrum  
method”.  ISET  Journal  of  Earthquake  Technology,  41:1-­‐13.  
 
Glaister,  S.  and  Pinho,  R.  (2003)  “Development  of  a  Simplified  Deformation-­‐Based  Method  for  
Seismic  Vulnerability  Assessment,”  Journal  of  Earthquake  Engineering,  Vol.  7,  pag  107-­‐
140.  
 
Goretti,   A.   and   Di   Pasquale,   G   (2004)   “Building   inspection   and   damage   data   for   the   2002  
Molise,  Italy  earthquake,”  Earthquake  Spectra,  Vol.  20,  pp.  S167-­‐S190  
 
Grunthal,   G.   (ed.),   (1998)   European   Macroseismic   Scale   1998   (EMS-­‐98).   Cahiers   du   Centre  
Europeen  de  Geodynamique  et  de  Seismologie  15,  Centre  Europeen  de  Geodynamique  
et  de  Seismologie,  Luxembourg.  
 
Hill,   M.   and   Rossetto,   T.   (2008a)   “Comparison   of   building   damage   scales   and   damage  
descriptions   for   use   in   earthquake   loss   modelling   in   Europe,”   Bulletin   of   Earthquake  
Engineering,  Vol.  6,  pp.  335-­‐365.  
 
Hill,   M.   and   Rossetto,   T.   (2008b)   “Improving   seismic   loss   estimation   for   Europe   through  
enhanced   relationships   between   building   damage   and   repair   costs,”   Proceedings   of  
14th  World  Conference  on  Earthquake  Engineering,  Beijing,  China,  Paper  S01-­‐02-­‐009.    
 
Hill,   M.   and   Rossetto,   T.   (2008c)   “Development   of   parameters   for   use   in   seismic   recovery  
estimation  of  residential  buildings  in  Lima,  Peru,”  Proceedings  of  14th  World  Conference  
on  Earthquake  Engineering,  Beijing,  China,  Paper  09-­‐01-­‐0027.  
 
Huyck,  C.,  Esquivias,  G.,  Gamba,  P.,  Hussain  M.,  Odhiambo,  O.,  Jaiswal,  K.,  Chen,  B.,  Becker,  M.,  
Yetman,   G.,   Crowley,   H.   (2011)   D2.2   Survey   of   available   input   databases   for   GED.  
Available  from  URL:  http://www.nexus.globalquakemodel.org/ged4gem/posts  
 
Jaiswal,   K.S.,   Wald,   D.J.   and   Hearne,   M.   (2008)   Estimating   Casualties   for   Large   Earthquakes  
Worldwide   using   an   Empirical   Approach,   U.S.   Geological   Survey   Open-­‐File   Report  

78
 

2009-­‐1136,  78  pp.  


 
Jaiswal,  K.S.,  and  Wald,  D.J.  (2008)  Creating  a  Global  Building  Inventory  for  Earthquake  Loss  
Assessment  and  Risk  Management,  U.S.  Geological  Survey  Open-­‐File  Report  2008-­‐1160,  
103  pp.  
 
Jaiswal,   K.S.   and   Wald,   D.J.   (2011)   Rapid   Estimation   of   the   Economic   Consequences   of   Global  
Earthquakes,  U.S.  Geological  Survey  Open-­‐File  Report  2011-­‐1116,  47  pp.  
 
Jayaram,   N.   and   Baker,   J.   W.   (2009)   “Correlation   model   for   spatially   distributed   ground-­‐
motion   intensities,”   Earthquake   Engineering   and   Structural   Dynamics,   38(15),   pp.  
1687–1708.  
 
Kappos,   A.,   Panagopoulos,   G.,   Panagiotopoulos,   C.,   Penelis,   G.   (2006).   "A   hybrid  
method   for   the   vulnerability   assessment   of   R/C   and   URM   buildings".   Bulletin  
of  Earthquake  Engineering,  4(4):391-­‐413.  
 
Karababa,   F.S.   and   Pomonis,   A.   (2010)   “Damage   data   analysis   and   vulnerability   estimation  
following   the   August   14,   2003   Lefkada   Island,   Greece   earthquake,”   Bulletin   of  
Earthquake  Engineering,  Vol.  9,  No.  4  
 
Lagomarsino  S.,  Giovinazzi,  S.  (2006).  “Macroseismic  and  Mechanical  Models  for  the  
Vulnerability   Assessment   of   Current   Buildings”.   Bulletin   of   Earthquake  
Engineering,  Special  Issue  “Risk-­‐Ue  Project”,  4(4).  
 
Liel,   A.B.   and   Lynch,   K.P.   (2009)   “Vulnerability   of   reinforced   concrete   frame   buildings   and  
their   occupants   in   the   2009   L’Aquila,   Italy   earthquake,”   Natural   Hazards   Review,   in  
press    
 
Lin,   Y.   Y.,   Chang,   K.   C.,   Wang,   Y.   L.   (2004).   “Comparison   of   displacement   coefficient  
method   and   capacity   spectrum   method   with   experimental   results   of   RC  
columns”.  Earthquake  Engineering  &  Structural  Dynamics,  33:35-­‐48.  
 
Masi,   A.,   Dolce,   M.   and   Vona,   M.   (2002)   A   procedure   to   estimate   economic   losses   due   to  
damage   of   residential   buildings   (in   Italian),   DiSGG   (Dipartimento   di   Strutture,  
Geotecnica,   Geologica   Applicata   all’Ingegneria)   Proceedings   5,   University   of   Basilicata,  
Potenza,  Italy.  
 
McCormack,   T.C.   and   Rad,   F.N.   (1997).   “An   earthquake   loss   estimation   methodology   for  
buildings  based  on  ATC-­‐13  and  ATC-­‐21”,  Earthquake  Spectra,  Vol.  13,  No.  4,  pp.  605-­‐
621.  
 
McGuire,   R.K.   (2004)   Seismic   Hazard   and   Risk   Analysis.   Earthquake   Engineering   Research  
Institute.  
 
MunichRe   (2012).   Topics   Geo:   Natural   catastrophes   2011   Analyses,   assessments,  
positions.    2012  Issue,  Munich,  Germany.  
 

79
Orsini   G.   (1999)   “A   model   for   buildings’   vulnerability   assessment   using   the   parameterless  
scale  of  seismic  intensity  (PSI),”  Earthquake  Spectra,  15(3),  pp.  463–83  
 
Park,   Y.J.   and   Ang,   A.H.S.   (1985).   “Mechanistic   seismic   damage   model   for   reinforced  
concrete”,  Journal  of  Structural  Engineering,  ASCE,  Vol.  111,  No.  4,  pp.  722-­‐739.  
 
Paulay,   T.,   Priestley,   M.   J.   N.   (1992).   Seismic   Design   of   Reinforced   Concrete   and  
Masonry  Buildings.  John  Wiley  and  Sons,  Inc.,  New  York,  USA.  
 
Pinho,  R.,  Bommer,  J.  J.,  Glaister,  S.  (2002).  “A  simplified  approach  to  displacement-­‐
based   earthquake   loss   estimation   analysis”.   Proceedings   of   the   12th   European  
Conference  on  Earthquake  Engineering,  London,  England.  
 
PRB   (2010).   World   population   data   sheet.   Population   Reference   Bureau,  
Washington  D.C.,  USA.  
 
Priestley,   M.   J.   N.,   Calvi,   G.   M.,   Kowalsky,   M.   J.   (2007).   Displacement-­‐based   Seismic  
Design  of  Structures.  IUSS  Press,  Pavia,  Italy.  
 
Reiter,  L.  (1990)  Earthquake  Hazard  Analysis:  Issues  and  Insights.  Columbia  University  Press,  
New  York.  
 
Rezaeian,   S.,   Bozorgnia,   Y.,   Idriss,   I.   M.,   Campbell,   K.,   Abrahamson,   N.,   Silva,   W.  
(2012).   Spectral   Damping   Scaling   Factors   for   Shallow   Crustal   Earthquakes   in  
Active   Tectonic   Regions.   PEER   Report   no   2012/01,   Pacific   Earthquake  
Engineering  Research  Center,  California,  USA.  
 
Rossetto,  T.  Elnashai,  A.  (2003)  “Derivation  of  vulnerability  functions  for  European-­‐type  RC  
structures   based   on   observational   data,”   Engineering   Structures,   Vol.   25,   pp.   1241-­‐
1263.  
 
Rossetto,   T.,   Elnashai,   A.   (2005)   “A   new   analytical   procedure   for   the   derivation   of  
displacement-­‐based   vulnerability   curves   for   populations   of   RC   structures”.  
Engineering  Structures  27(3):397-­‐409  
 
Rota,   M.,   Penna,   A.,   Strobbia,   C.L.   (2008)   “Processing   Italian   damage   data   to   derive  
typological   fragility   curves,”   Soil   Dynamics   and   Earthquake   Engineering,   Vol.   28,   No.  
10-­‐11,  pp.  933-­‐947  
 
Salvatore,   M.,   Pozzi,   F.,   Ataman,   E.,   Huddleston,   B.   and   Bloise,   M.   (2005)   “Mapping   global  
urban   and   rural   population   distributions,”   Food   and   Agriculture   Organization   of   the  
United   Nations,   Rome,   2005.   Available   from   URL:  
http://www.fao.org/docrep/009/a0310e/A0310E00.htm#TOC  
 
Sarabandi,   P.,   Pachakis,   D.,   King,   S.,   Kiremidjian,   A.   (2004)   “Empirical   fragility   functions   from  
recent   earthquakes,”   Proceedings   of   13th   World   Conference   on   Earthquake   Engineering,  
Vancouver,  Canada.    
 

80
 

Silva,  V.,  Crowley,  H.,  Pinho,  R.,  Varum,  H.  (2013)  “Extending  Displacement-­‐Based  Earthquake  
Loss   Assessment   (DBELA)   for   the   Computation   of   Fragility   Curves”.   Engineering  
Structures,    56:343-­‐356.  
 
Silva,   V.,   Crowley,   H.,   Pinho,   R.,   Varum,   H.   (2014)   “Seismic   Risk   Assessment   for   mainland  
Portugal”.  Bulletin  of  Earthquake  Engineering,  13(2):429-­‐457.  
 
Silva,   V.,   Crowley,   H.,   Pinho,   R.,   Varum,   H.   (2014)   “Investigation   of   the   Characteristics   of  
Portuguese  Regular  Moment-­‐frame  RC  Buildings  and  Development  of  a  Vulnerability  
Model”.  Bulletin  of  Earthquake  Engineering,  DOI  10.1007/s10518-­‐014-­‐9669-­‐y.  
 
Silva,   V.,   Crowley,   H.,   Pagani,   M.,   Monelli,   D.,   Pinho,   R.   (2014)   “Development   of   the  
OpenQuake   engine,   the   Global   Earthquake   Model’s   open-­‐source   software   for   seismic  
risk  assessment”.  Natural  Hazards,  72(3):1409-­‐1427.  
 
Singhal,   A.   and   Kiremidjian,   A.S.   (1996)   “Method   for   probabilistic   evaluation   of   seismic  
structural  damage,”  Journal  of  Structural  Engineering,  ASCE,  Vol.  122,  No.  12,  pp.  1459-­‐
1467  
 
So,   E.   (2009)   The   assessment   of   casualties   for   earthquake   loss   estimation,   PhD   Thesis,  
University  of  Cambridge,  UK.  
 
Spence,   R.   Ed   (2007)   Earthquake   disaster   scenario   predictions   and   loss   modelling   for   urban  
areas,  LESSLOSS  Report  7,  IUSS  Press,  Pavia,  Italy.  
 
UNISDR   (2009).   “Global   Assessment   Report   on   Disaster   Risk   Reduction   2009”.  
Report,   United   Nations   International   Strategy   for   Disaster   Reduction  
Secretariat,  Geneva,  Switzerland.  
 
Vamvatsikos,   D.,   Cornell,   A.   C.   (2002).   “Incremental   dynamic   analysis”.   Earthquake  
Engineering  &  Structural  Dynamics,  31(3):491-­‐514.  
 
Vamvatsikos   D,   Cornell   CA   (2006)   Direct   estimation   of   the   seismic   demand   and   capacity   of  
oscillators   with   multilinear   static   pushovers   through   Incremental   Dynamic   Analysis,  
earthquake  Engineering  and  Structural  Dynamics  35(9):1097–1117.  
 
Wald,   D.,   Quitoriano,   V.,   Heaton,   T.,   Kanamori,   H.,   Scrivner,   C.,   Worden,   B.   (1999)   "TriNet  
ShakeMaps:   Rapid   generation   of   peak   ground   motion   and   intensity   maps   for  
earthquakes  in  southern  California,"  Earthquake  Spectra,  Vol.  15  (N°  3),  pp.  537-­‐555.  
 
Wasserman,   L.   (2004)   All   of   Statistics:   A   Concise   Course   on   Statistical   Inference.  
Springer,  New  York,  USA.  
 
Whitman,  R.V.,  Reed,  J.W.  and  Hong,  S.T.  (1973)  “Earthquake  Damage  Probability  Matrices,”  
Proceedings  of  5th  World  Conference  on  Earthquake  Engineering.  

81

You might also like