You are on page 1of 604

National Science Geotechnical Extreme Applied Technology

Foundation Events Reconnaissance Council

Sissy Nikolaou (WSP | Parsons Brinckerhoff)


Xavier Vera-Grunauer (Universidad Católica de Guayaquil, Geoestudios)
Sissy Nikolaou (WSP | Parsons Brinckerhoff)
Xavier Vera-Grunauer (Universidad Católica de Guayaquil, Geoestudios)
Ramon Gilsanz (Gilsanz Murray Steficek, LLP)

Universities Army Corps of Engineers


BERKELEY Roberto Luque, Tadahiro Kishida ECACE Enrique Morales Moncayo
BYU Kyle Rollins
CORNELL Tom O’Rourke, Bernardo Casares Universities
RPI Nonika Antonaki, Panagiota Kokkali ESPE Theofilos Toulkeridis
STANFORD Eduardo Miranda IG-EPN Hugo Yepes
UARK Clinton Wood Alexandra Alvarado
UBUFFALO Enrique Morales, Jerome S. O'Connor SFQUITOU Fabricio Yepez
Andreas Stavridis UCSG Adolfo Caicedo
UMICH Adda Athanasopoulos-Zekkos Francisco Ripalda
Oscar González Xavier Vera-Grunauer
Agencies and Individual Volunteers Engineering Firms
USACE Gabriela M. Lyvers Geoestudios Xavier Vera-Granauer
FHWA Daniel Alzamora Noemí Villagrán León
Individuals Pablo López, Said Maalouf Ignacio Ochoa, Danilo Davila
Engineering Firms Droneando Luis Fernando Rodríguez

GMS Ramon Gilsanz, Virginia Diaz,


Laura Hernandez
WSP | PB Sissy Nikolaou, Guillermo Diaz-Fanas UNCO Carlos Arteta
Nonika Antonaki, Efthymios Vaxevanis
Panagiota Kokkali, Jay Mezher
Patrick Bassal

WSP | PB Guillermo Diaz-Fanas (for GEER-Ecuador)


GMS Virginia Diaz (for ATC-Ecuador)
BERKELEY Fernando "Estefan" Garcia (2016 GEER Recorder)
Blank Page
DEDICATION

To the resilient people of Ecuador


Blank Page
Acknowledgements
The information contained in this report is based upon work supported by the National
Science Foundation through the Geotechnical Engineering Program under Grant No. CMMI-
CMMI-1266418. Any opinions, findings, and conclusions or recommendations expressed in
this material are those of the authors and do not necessarily reflect the views of the NSF.

The GEER Association is made possible by the vision and support of the NSF Geotechnical
Engineering Program Directors: Dr. Richard Fragaszy and the late Dr. Cliff Astill. GEER
members donate their time, talent, and resources to collect time-sensitive field observations of
the effects of extreme events. The GEER Association web site, which contains additional
information, may be found at: http://www.geerassociation.org.

Financial contribution from the Applied Technology Council (ATC) funded the
participation of two structural engineers in the mission. US agencies, academic institutions,
and individuals volunteered with time and/or resources, gratefully acknowledged in Chapter 3.

The engineering firms of WSP | Parsons Brinckerhoff (WSP | PB) and Gilsanz, Murray,
Steficek, LLP (GMS) provided staff time of 18 engineers and resources, and participated in
data reduction, report preparation and related presentations. The Geotechnical and Tunneling
Technical Excellence Center (TEC) of WSP | PB volunteered time and shared technology from
two of their core developmemt initiatives on Multihazards Resilience and Virtual Design &
Construction in the data and 3D mapping modeling from drone and camera images.

The Ecuadorian government agencies and army, academic institutions, private firms, and
individuals were instrumental in the success of this mission. Public agencies included the
Ministries of Urban Development & Housing (MIDUVI) and of Transportation & Public
Works, the National Geophysical Institute and the 911 Emergency Response Agency. The
Army Corps of Engineers (ECACE) arranged for helicopter overflies and provided data, maps
and statistical and design information. Senior faculty and students from the major universities
ESPE, ESPOL, IG-EPN, SFQUITOU, and UCSG with the Colombian Univ. Norte participated
and gratefully listed in Chapter 3 with several individuals. Engineering firms worked closely
with the US reconnaissance team, performed in-situ and soil laboratory and structural material
testing, and provided input for the report. Listed with many thanks in Chapter 3, they are
GEOESTUDIOS, Droneando, UNCO, CVA, CPR, Nylic, Constructora Verdu, and AET.

GEER-ATC Muisne, Ecuador, Earthquake Acknowledgements


Report Version 1 i
Blank Page
TABLE OF CONTENTS

Page

Dedication

Acknowledgements i

Table of Contents ii

List of Figures iv

List of Tables xxxv

Chapter Page

1 INTRODUCTION 1-1

2 ECUADOR AND RECONNAISANCE 2-1

2.1 Ecuador 2-1


2.2 Reconnaissance Focus 2-9
2.3 Reconnaissance Routes and Objectives 2-11

3 GEER – ATC TEAM 3-1

4 SEISMOTECTONICS, SEISMIC HISTORY 4-1

4.1 Seismotectonic Setting 4-1


4.2 Historic Seismicity 4-11

5 STRONG GROUND MOTIONS 5-1

5.1 Seismological Aspects 5-1


5.2 Strong Motion Records 5-5
5.3 Tsunamis 5-48

6 RECONNAISSANCE METHODS 6-1

6.1 Geotechnical Exploration Methods 6-1


6.2 Drones 6-10
6.3 Helicopter Flyovers 6-13
6.4 Security Cameras 6-13

7 GEOTECHNICAL OBSERVATIONS 7-1

7.1 Geologic Setting 7-1

GEER-ATC Muisne, Ecuador, Earthquake Table of Contents


Report Version 1 ii
7.2 Site Characterization Data 7-5
7.3 Site Effects 7-41
7.4 Liquefaction 7-59
7.5 Earth Retaining Structures 7-79
7.6 Landslides and Rock Falls 7-94
7.7 Transportation Network 7-110
7.8 Dams 7-139
7.9 Embankments 7-163
7.10 Case Studies 7-174

8 STRUCTURAL OBSERVATIONS 8-1

8.1 Building Inventory and Construction Types 8-1


8.2 Construction Means/Methods/Implementation 8-6
8.3 Building Codes and Evolution 8-14
8.4 Damage Assessment/Tagging/Incorporation of ATC-20 8-21
8.5 Main Structural Observations and Structural Behavior 8-28
8.6 Non-structural Components 8-40
8.7 Residential and Commercial Buildings 8-55
8.8 Essential/Critical Buildings 8-61
8.9 Special Structures, Tanks, Cmaroneras 8-76
8.10 Case Studies 8-76

9 COMMUNITY PREPAREDNESS & RESPONSE 9-1

9.1 Introduction and the Red Cross 9-1


9.2 Army Action Plan and Immediate Response 9-2
9.3 Zone Zero 9-8
9.4 Homeless Relocation 9-2
9.5 Supplies and Medical Services 9-11
9.6 Fuel and Fires 9-19

10 CONCLUSIONS 10-1

11 REFERENCES 11-1

APPENDIX A PRE-EARTHQUAKE GEOTECHNICAL INVESTIGATIONS A-1

APPENDIX B POST-EARTHQUAKE GEOTECHNICAL INVESTIGATIONS B-1

GEER-ATC Muisne, Ecuador, Earthquake Table of Contents


Report Version 1 iii
LIST OF FIGURES
Figure Page

2.1.1 Ecuador Geography: Location in South America. Base image: World Atlas 2-1
(©2016).

2.1.2 Ecuador provinces (top, Wikipedia, 2016) and regions (Your Escape to Ecuador, 2016). 2-2
(Your Escape to Ecuador, 2016).

2.1.3 Quito. Base image: Lifewithoutlimbs (©2013). 2-4

2.1.4 City of Guayaquil. Base image: Municipal GAD Guayaquil (©2014). 2-4

2.1.5 Sea turtles off the coast of the Galápagos Islands (Avalon Waterways, ©2016) 2-6

2.1.6 Sunset at Galápagos Islands in the Pacific Ocean (Green Coconut Rum, ©2016). 2-7

2.1.7 Republic of Ecuador National Assembly. Base image: La Info ( ©2014). 2-8

2.2.1 GEER-ATC team reconnaissance on the ground with focus cities shown with red pins. 2-9
Dams observed by military helicopter flyover shown as white pins. Base Image: Google
Earth (©2016).

2.3.1 Team 1, 4/27/16: GPS track for measurements at Manta strong motion station and port. 2-11
Base Image: Google Earth (©2016).

2.3.2 Teams 2, 3, 4/27/16: GPS track at Manta Port. Base Image: Google Earth (©2016). 2-12

2.3.3 Team 4, 4/27/16: GPS track of structural damage survey at Manta. Base Image: 2-12
Google Earth (©2016).

2.3.4 Team 1, 4/28/16: GPS track to observe behavior of bridges in Rocafuerte and 2-13
Portoviejo. Base Image: Google Earth (©2016).

2.3.5 Teams 2, 3, 4/28/16: GPS tracks for essential buildings. Base Image: G. Earth ( ©2016). 2-13

2.3.6 Teams 1, 2, 4/29/16: GPS route to Pedernales, Base Image: Google Earth ( ©2016). 2-14

2.3.7 Team 1, 4/29/16: GPS survey route in Jama. Base Image: Google Earth ( ©2016). 2-14

2.3.8 Team 1, 4/29/16: GPS survey route in Pedernales. Base Image: Google Earth (©2016). 2-15

2.3.9 Team 2, 4/29/16: GPS survey route of Los Perales Airport in San Vicente. Base 2-15
Image: Google Earth (©2016).

2.3.10 Team 2, 4/29/16: GPS survey route in Canoa. Base Image: Google Earth ( ©2016). 2-16

2.3.11 Team 2, 4/29/16: GPS survey route in Pedernales. Base Image: Google Earth (©2016). 2-16

2.3.12 GEER-ATC Team, 4/30/16: Helicopter paths Base Image: Google Earth (©2016). 2-17

2.3.13 Team 1, 4/30/16: GPS track of Vs surveys at Chone dam. Base Image: G. Earth (©2016). 2-17

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 iv
2.3.14 Teams 2, 3, 5/1/16: GPS locations of Vs surveys at Manta ground zero. Base Image: 2-18
Google Earth (©2016).

4.1.1 The geodynamic setting of Ecuador and associated plates, microplate and 4-2
volcanic ridges (Oceanic Plates: Pacific, Nazca, Cocos; Continental Plates: Caribbean
and South American; Oceanic Microplate: Galapagos, Ridges: Carnegie, Cocos (after
Toulkeridis, 2013)

4.1.2 Geodynamic setting of Ecuador with main cities shown with white letters: Pa=Pasto; 4-3
Q=Quito; R=Riobamba; B=Bahía; G=Guayaquil; Cu=Cuenca; Ta=Talara;
Bu=Buenaventura. Base map modified from GeoMapApp (geomapapp.org).

4.1.3 Ocean floor ages along South America’s Nazca plate with major Fracture . Zones (FZ) 4-4
and Ridges (R), and locations of 6 subduction earthquakes of M w greater than 8.3 since
1900 with epicenter (in white circle) near a FZ or R and focal plane shown where
available. JFR = Juan Fernandez R, IQR = Iquique R, PA = Pacific, NZ = Nazca,
SA = South America (Carena, S., 2011 and Müller et al., 1997).

4.1.4 The Galapagos spreading center is an EW ridge rising above the seafloor with 4-5
shallowest depth ~1.6 km north of the Islands. Areas deeper than 3.5 km are shown in
blue (Haymon et al., 1993).

4.1.5 Subduction zone of Ecuador and its volcanic chains (after Toulkeridis, 2013). 4-6

4.1.6 Left: Guagua Pichincha (0°10'14.88"S, 78°36'45.36"W), 10/7/99; U. right: Tungurahua, 4-7
2002 (1°28'12.71"S, 78°26'41.28"W); Center: Sangay, 2001 (2°00'17.99"S, 78°20'26.88"
W); L. right: Reventador, 2002 (0°04'39.00"S, 77°39'20.87"W). Photos by A. Speck &
Anonymous (Reventador).

4.1.7 Volcanic Explosivity Index (VEI) scale and ejecta volume (volcanoes.usgs.gov). 4-7

4.1.8 Carnegie Ridge in 3-D imaging. 4-9

4.1.9 Tsunamis epicenters as red points during the last century (Toulkeridis, 2013). 4-10

4.1.10 Photo of damage after the 1979 tsunami in Colombia (credit: El Pais, Colombia). 4-11

4.2.1 Intensity database for Ecuador (Egred, 2009), containing events with at least one 4-12
intensity VI reported (from Beauval et al. 2013).
4.2.2 Epicenters from the unified earthquake catalog 1587–2009, integrating instrumental 4-13
and historical earthquakes, displaying magnitude 4 and above (after Beauval et al., 2013).

4.2.3 Current and historical seismicity related to the 2016 earthquake rupture zone (Beauval 4-14
et al., 2013). Main cities: Q=Quito; M=Manta; P=Portoviejo; G=Guayaquil.

4.2.4 Image from minimum ISC model by Chlieh et al. (2014). The 2016 megathrust 4-15
earthquake almost closes the gap that remained unbroken since the 1906 rupture.

4.2.5 Ecuador historic seismicity (modified from Swenson and Beck, 1996). 4-15

4.2.6 The 1797 Riobamba earthquake: Isoseismal curves extrapolated from observed 4-18
local intensities, coseismic fissures and cracks, and landslides (Egred, 2000; Baize et
al., 2015).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 v
4.2.7 1906 Ecuador earthquake photo (Enciclopedia Encarta, Library of Congress). 4-19

4.2.8 Actual records of G2 and G3 waves from the 1906 Colombia-Ecuador earthquake by 4-20
a NS-component Wiechert seismograph at Göttingen, Germany (Kanamori & McNally,
1982).
4.2.9 Topography affected by the 1949 Ambato earthquake (Life magazine, Aug 22, 1949). 4-21

4.2.10 Ambato hospital ruined by the 1949 earthquake (photo by G.E. Lewis, USGS library). 4-21

4.2.11 Ambato Cathedral rescuers work to dig out bodies of 70 people, including 60 children, 4-22
who sought shelter in the centuries-old church (photo by Life magazine, 1949).

4.2.12 Fissures on the ground surface following the 1949 Ambato earthquake (Life, 1949). 4-22

4.2.13 The village of Pelileo flattened by the Ambato earthquake (Life magazine, 1949). 4-23

4.2.14 Rayleigh waves recorded by a Press-Ewing seismograph at Pasadena during the 1958 4-24
and 1979 events, and used to derive the M w of the 1958 event (Kanamori & McNally,
1982).

4.2.15 1979 Ecuador-Colombia earthquake and tsunami (photos by G. Pararas-Carayannis). 4-24

4.2.16 Record of 1979 tsunami as recorded in Esmeraldas (Pararas-Carayannis, 1980). 4-25

4.2.17 Tsunami heights from the 1906 and 1979 earthquakes as they arrived in tide stations 4-25
(left; Kanamori & McNally, 1982); refraction diagram and travel time for the 1979
tsunami across the Pacific Ocean (right; S. Poole, in Pararas-Carayannis, 1980).

4.2.18 Looking downstream at the confluence of the Río Malo and the Río Coca, NE Ecuador. 4-26
Both channels have been choked by debris flows triggered by the 1987 earthquakes
(USGS, 2001).

4.2.19 Valley wall of Río Malo, NE Ecuador, after the 1987 earthquakes with denudation of 4-27
slopes due to avalanches/flows and of valley bottom due to debris flows and flooding
(USGS, 2001).

4.2.20 Route and station locations of the Trans-Ecuadorian pipeline (top) and area of major 4-27
damage (Crespo et al., 1987).

4.2.21 Trans-Ecuadorian pipeline: damage at east bank of the Aguarico River (left); view from 4-28
west bank with pipe pulled south by flood with remnant bridge pier and abutment
(Crespo et al., 1991).

4.2.22 Building Calypso, before and after the Bahía de Caráquez earthquake (Castro, 2007). 4-29

5.1.1 Focal mechanism for the main 4/16/216 event reported in the Global CMT catalog 5-1
(Dziewonski et al., 1981; Ekström, G. et al., 2012).

5.1.2 Epicenters of the main event (purple star) and its aftershocks (circles colored and sized 5-2
according to magnitudes). Epicenters of historical earthquakes are shown in yellow stars
and background seismicity prior to April 16th is shown as black and white circles.

5.1.3 The 110 years of instrumental seismicity in Ecuador, 1900–2009. Earthquakes with 5-3
Mw ≥7 are plotted as stars (Yepez et al., 2016).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 vi
5.2.1 Spatial distribution of the RENACand OCP accelerometer stations. The epicenter of 5-5
the main event is shown as a star, and the circles at selected stations reflects the Peak
Ground Acceleration (PGA) recorded in m/s2 (by IG-EPN, Singaucho et al., 2016).

5.2.2 The two main types of strong motion instruments Guralp CMG-5TDE (left) and Reftek 5-7
130-SMA (right). Details can be found in EG-IPN (2016s).

5.2.3 Contours of Intensity (ShakeMap) from USGS and Strong Motion Stations (SMS), 5-9
color-coded by the intensity of the motion in terms of the geo-mean Peak Ground
Acceleration (PGA).

5.2.4 Acceleration time series in the EW component overlying a map of Ecuador with main 5-9
cities and recording stations. Colored circles same as in Fig. 5.2.3.

5.2.5 Acceleration, velocity and displacement time histories for the EW component of 5-10
stations APED, AMNT, ACHN and APO1.

5.2.6 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AAM2, Ambato. 5-11

5.2.7 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ACH1, Machala. 5-12

5.2.8 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ACHN, Chone 5-13

5.2.9 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ACUE, Cuenca. 5-14

5.2.10 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station PDNS, 5-15
Pedernales.

5.2.11 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station APED, 5-16
Pedernales.

5.2.12 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AV18, Quininde. 5-17

5.2.13 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AV21, Viche. 5-18

5.2.14 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AMA1, 5-19
Esmeraldas.

5.2.15 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AES2, 5-20
Esmeraldas.

5.2.16 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station APR2, Puerto 5-21
Quito.

5.2.17 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ASDO, Santo 5-22
Domingo.

5.2.18 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ALOR, San 5-23
Lorenzo.

5.2.19 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station APO1, Portoviejo. 5-24

5.2.20 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AMNT, Manta. 5-25

5.2.21 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station EPNL, Quito. 5-26

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 vii
5.2.22 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AV11, 5-27
Guallabamba.

5.2.23 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AOTA, 5-28
Otavalo.

5.2.24 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AIB1, Ibarra. 5-29

5.2.25 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AIB2, Ibarra. 5-30

5.2.26 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ALAT, 5-31
Latacunga.

5.2.27 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station APS4, Papallacta. 5-32

5.2.28 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ATUL, Tulcán. 5-33

5.2.29 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AGYE, 5-34
Guayaquil.

5.2.30 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AMIL, Milagro. 5-35

5.2.31 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ALIB, 5-36
La Libertad.

5.2.32 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station ALG1, Loja. 5-37

5.2.33 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station GYE1, Guayaquil. 5-38

5.2.34 Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AGYE, 5- 39
Guayaquil.

5.2.35 Observed ground motions (green dots), predicted by the BCHydro GMPE median (red 5-41
and median curve for main Ecuador event (blue
dashed line).

5.2.36 Between-event residuals for PGA showing points used for the BCHydro GMPE 5-42
regression with Maule, Chile (2010), Tohoku, Japan (2011), and Muisne, Ecuador
(2016) earthquakes.

5.2.37 Within-event residuals (δWes) for PGA and Sa at periods of 0.2, 1.0 and 3.0 s. 5-42

5.2.38 Comparison of GMPE predicted 5% damped median and +/- 1 standard deviation 5-43
response spectra (red lines) with recorded geo-mean response spectra (blue lines) for
stations (a) APED, (b) PDNS, (c) AMNT, (d) APO1, (e) ACHN, (f) AV21, (g) AGYE
and (h) AGY2.

5.2.39 Map of NEC-15 “Z” factor, equal to the PGA in rock or stiff soil conditions, for a 5-45
return period of 475 years or 10% probability of exceedance in 50 years (NEC, 2015).

5.2.40 Shape of design elastic acceleration spectra from NEC-15. 5-45

5.2.41 Comparison of recorded geometric-mean, median GMPE predictions, NEC-15 design 5-47
spectra for stations (a) APED, (b) PDNS, (c) AMNT, (d) APO1, (e) ACHN, (f) AGY2.

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 viii
5.3.1 Tsunami water levels from the main earthquake registered at NOAA 32411 (top) 5-48
and 32413 (bottom) DART buoys in the Panama and Peru basins, respectively.
Information by NOAA / PMEL / Center for Tsunami Research.

5.3.2 Tsunami water level during the main earthquake registered at the local 5-49
INOCAR-DART buoy which is close to the epicenter (INOCAR, 2016).

6.1.1 Standard Penetration Test: Driving split-barrel sampler (Mayne et al., 2002). 6-1

6.1.2 Cone Penetration Test: Procedure and components (Mayne et al., 2002). 6-2

6.1.3 Surface wave method: from field acquisition to final Vs profile (C. Wood, 6-3
U. Arkansas).

6.1.4 MASW process from field acquisition to final Vs profile (C. Wood, U. Arkansas). 6-4

6.1.5 MAM process from field acquisition to final Vs profile (C. Wood, U. Arkansas). 6-5

6.1.6 Reconnaissance MASW testing setup with GEER team members K. Rollins, 6-6
A. Caicedo, C. Wood working on-site (GPS: 0o42'1.6''S, 79o59'21.7''W).

6.1.7 Reconnaissance active MAM field testing setup. 6-7

6.1.8 Reconnaissance HVSR setup and components. 6-8

6.2.1 Reconnaissance drones: Phantom 4 (left, DJI, 2016a); Inspire 1 Pro (right, 6-10
DJI, 2016b).

6.2.2 3D mapping examples using Pix4D software: (a) Manta; (b) Potrovejo 6-10

6.2.3 Orthomap example taken at an altitude of 37 m. 6-11

6.2.4 3D modeling developed by team member L. Rodriguez for: (a) IEES Manta 6-12
Hospital (~GPS 0°57'18.2"S, 80°43'27.0"W); (b) landslide near San Isidro
(~GPS 0°25'0.6"S, 0°15'40.1"W); and (c) church in Montecristi, Manabi
(~GPS 1°3'0.0''S, 80°40'0.0''W).

6.3.1 Reconnaissance team members before boarding the Ecuadorian Army helicopter 6-13
at the Portovejo army base. From left to right: R. Gilsanz, V. Diaz, D. Alzamora,
S. Nikolaou, C. Wood, X. Vera, K. Rollins, A. Caicedo, R. Luque, G. Lyvers,
G. Ponte Vasquez (GPS 1° 3' 7"S, 80° 28' 30"W).

6.3.2 GPS track of reconnaissance helicopter flight provided by the Ecuadorian Army 6-13

6.3.3 Ecuadorian air force helicopter (bot.), pilot (upper left corner) and crew officers 6-14
(mid. left and 3rd right row). General Pedro Mosqueda of the EC Army Corps of
Engineers is discussing the rules of the flight and the routes with reconnaissance
team members (2nd right row).

6.3.4 Highway E15 flyover observations: stable steep slopes (left) and landslide (right). (GPS 6-15
left: 0° 7' 15.86" S, 80° 13' 15.28" W; GPS right: 0°33' 5"S, 80°25'42"W).

6.3.5 Chone Dam flyover observations (GPS 0° 51' 58.3"S, 80° 6' 37.6" W). 6-15

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 ix
6.3.6 Manta port from flyover at altitude of 35.7 m (GPS 0° 56' 29" S, 80° 43' 20" W). 6-16

6.3.7 Camaroneras farms from an altitude of 92.9 m (top, GPS 0° 37' 38.4" S, 6-16
80° 22' 47.9"W). Liquefaction evidence near Bahia and S. Vicente (bot. left) and slope
failure near Pedernales (bot. right).

6.4.1 Integrated Security System ECU 911 operation center to monitor emergencies 6-17
in Ecuador.

6.4.2 Services that ECU 911 provides to citizens include (clockwise from top left): 6-18
video surveillance; emergency phone lines; community outreach; and
institutional coordination.

6.4.3 Images from camera GYE 61F1 near downtown Guayaquil (9 de Octubre and 6-18
Padre Solano Streets): (a) before, and (b) during the main seismic event.
Orthomap example taken at an altitude of 37 m.

6.4.4 Camera GYE 157 D images (Chile & 9 de Octubre Streets), before, during, and 6-19
after the event.

6.4.5 Images from camera GYE 146D in Guayaquil (9 de Octubre & P. Carbo St.) 6-19

6.4.6 Camera Manta23 images from the avenue before (upper left) and after the event. 6-20

6.4.7 Camera Manta31 images showing Barrio Altamira, before (upper left) and after 6-20
after the event.

6.4.8 Images from SAM07F at la Puntilla (3.3-km), before (top) and during the event 6-21

6.4.9 Camera Porto09 images captured Manabí and Atahualpa St. near Imperial Hotel 6-21
before (top), during (middle) and after the seismic event (bottom).

6.4.10 Camera images from an office in Canoe (youtube.com/watch?v=Nz_05mT-Me0) 6-22

6.5.1 GEER-ATC members taking reconnaissance photos: V. Diaz and G. Lyvers, 6-23
Portovejo (left, 1o 3' 16.5'' N, GPS: 80o 26' 42.3'' E); C. Wood and others, Manta
(top right, GPS 0o 56' 28.3'' N, 80o 44' 4.1'' E); R. Luque, helicopter 83 m above
Los Cuyeyes (bot. right, GPS 1o 6' 57.7'' N, 80o 11' 29.5'' E).

6.5.2 Manta strong motion accelerometer station with coordinates identified by smart 6-24
phone. On the photo: G. P. Vasquez (GPS: 0o 56' 28'' S, 80o 44' 3'' W, El. 33 m).

6.5.3 Handheld GPS tracking device types used in reconnaissance (Garmin MAP®64 6-24
tracker - middle; Forerunner 305 Running watch - top right; Foretrex 401 Hiking
watch - bottom right). One member of each team deployed had a tracking device
(C. Wood on the left, Manta, GPS 0o 56' 29.9'' N, 80o 43' 57.1'' E). Additional
devices were coordinated for accuracy at the start of each mission.

6.5.4 GEER-ATC reconnaissance survey routes on April 29, 2016 in Pedernales, 6-25
Manabí: team 1 (left) and team 2 (right). Base Image by Google Earth (©2016).

6.5.5 Manta port deformations along the interface between quay wall & sidewalk 6-26
(GPS 0o 56' 13.2'' N, 80o 43' 27.4'' E). Measurements using metallic tape (bot.
left) by A. Caicedo (top, right) and visual evaluation of deformations in areas

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 x
hard to reach by D. Alzamora and A. Zekkos (bottom, right). Laser-based
equipment (center) were used to measure distances (left) and tilts (right).

6.5.6 Manta port wall fissure (left, K. Rollins, 0o 56' 29.6'' N, 80o 43' 27.4'' E); 6-27
Portovejo cemetery (top right, A. Caicedo, 1o 3' 14.6'' N, 80o 26' 52.1'' S); Manta
R. Zambrano Hospital bottles (bot., G. Lyvers, 0o 57' 15.4'' N, 80o 44' 32.8'' E).

6.5.7 Manta port parking lot wall (GPS 0o 56' 24.3'' N, 80o 43' 32.7'' E) tilting 6-27
measurements using rigid tape (A. Caicedo, left) and smartphone app
(A. Zekkos and K. Rollins, right).

6.5.8 Soil deformations and pavement failures between monuments at the Portovejo 6-28
cemetery measured by team member G. Lyvers (1o 3' 14.4'' N, 80o 26' 49.2'' E).

7.1.1 Geo-structural domains of Ecuador (Litherland and Zamora, 1991; 7-1


Bolaños, 2010).

7.1.2 Geologic map of Ecuador, northern Peru, and southern Colombia (after 7-2
Spikings et al. 2000) with fault notations; Banos (BF), Cosanga (CF), lngapirca
(IN), Llanganates (LF), Peltetec (PF).

7.1.3 Bedrock Geology. Sedimentary rock formations on the coastal west and igneous 7-3
and metamorphic rock formations at the mountainous areas inland (USGS, 1997).

7.1.4 Bedrock formation K-Ar ages from Jurassic to Early Cretaceous metamorphic 7-3
rocks and Tertiary plutons of the Cordillera Real (Litherland et al., 1994).

7.1.5 Mesozoic evolution of Ecuador (Kennerley, 1980; Brookfield et al., 2009). 7-4

7.2.1 Site locations with pre- and post-earthquake geotechnical data 7-5
(Google Maps ©2016).

7.2.2 Locations of 6 pre-earthquake projects with geotechnical data 7-7


(Google Maps ©2016).

7.2.3 Port of Manta (GPS Coordinates: 0°57'34.9992"S, 80°43'1.9914"W). Photo by 7-8


LCA (2016).

7.2.4 Manta Port: Pre-earthquake boring locations (Google Maps ©2016). 7-8

7.2.5 View of the Mejia Bridge prior to the earthquake (Google Maps ©2016). 7-10

7.2.6 Mejia Bridge: Geotechnical test locations pre-earthquake (yellow squares) and 7-11
post-earthquake. (Google Maps ©2016).

7.2.7 Los Caras Bridge (Flyg London, 2011). 7-12

7.2.8 Las Chacras Bridge: Pre-earthquake photo. Base image: CNES/Astrium (©2016). 7-12

7.2.9 Las Chacras Bridge: Location of pre-earthquake borings (Google Maps 7-13
©2016).

7.2.10 Boca de Briceno Bridge aerial photo (GPS Coordinates: 7-14


0⁰30'58.27''S, 80⁰26'29.08''W).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xi
7.2.11 Locations of pre-earthquake borings at the Briceño Bridge (adapted from Google 7-14
Maps© 2016).

7.2.12 Chone/Rio Grande Dam aerial view (GPS: 0°41'44.1''S, 79°59'30.6''W). 7-15

7.2.13 Locations of pre-earthquake borings at the Chone Dam (adapted from Google 7-16
Maps© 2016).

7.2.14 Locations of 20 post-earthquake GEER investigations (Google Maps ©2016). 7-17

7.2.15 Manta Port parking area before the earthquake with locations of active and 7-18
passive surface wave arrays, HVSR. Base image: Digital Globe (©2016).

7.2.16 Manta Port parking area post-earthquake: GEER team member C. Wood at the 7-19
active MASW array (GPS: 0° 56' 25.9" S, 80° 43' 31.8" W).

7.2.17 Manta Port parking area: Median theoretical Rayleigh wave dispersion curve 7-20
with fit to experimental dispersion data (left). Shear wave velocity Vs profile
from inversion (right).

7.2.18 Manta Port parking area: Mean horizontal-to-vertical spectral ratio (HVSR) 7-20
obtained. The red lines are the theoretical H/V spectral ratio curve and peak for
the developed Vs profile.

7.2.19 Manta Port wharf deck before the earthquake with locations of active surface 7-21
wave arrays. Base image: Digital Globe (©2016).

7.2.20 Manta Port wharf deck post-earthquake: GEER team member C. Wood with 7-21
an Ecuadorian colleague at the active MASW array (GPS: 0° 56' 14.9" S,
80° 43' 28.6" W).

7.2.21 Manta Port wharf deck: Median Rayleigh wave dispersion curve; fit to 7-22
experimental data (left). Median of top 1,000 Vs profiles from > 500,000
models in inversion search (right).

7.2.22 Manta Port: MAM and MASW Vs at Sites (Stations) 1, 2, 3 (GPS: 0° 55' 58.8" S, 7-22
80° 43' 19.2" W, 0° 55' 58.8" S, 80° 43' 22.8" W, 0° 56' 6" S, 80° 43' 22.8" W).

7.2.23 Mejia Bridge MAM and MASW Vs at Sites 1, 2 (GPS: 0° 59' 27.6" S, 80° 28' 8.4" W, 7-23
0° 59' 24.0" S, 80° 28' 12.0" W).

7.2.24 Las Chacras (Rio Chico) Bridge before the earthquake with locations of the active 7-24
surface wave array. Base image: CNES/Astrium (©2016).

7.2.25 Las Chacras (Rio Chico) Bridge post-earthquake: GEER team member A. Zekkos 7-24
with an Ecuadorian colleague at the active MASW array (GPS: 0° 58' 34.4" S,
80° 25' 20.0" W).

7.2.26 Las Chacras (Rio Chico) Bridge: Median theoretical Rayleigh wave dispersion curve; fit 7-25
to experimental dispersion data (left). Vs profile obtained from during inversion (right).

7.2.27 Chone Dam post-earthquake photo showing the locations of active and passive surface 7-26
wave arrays, and HVSR (GPS: 0° 41' 41.7" S, 79° 59' 21.9" S).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xii
7.2.28 Chone Dam: Active MASW array at top of dam (GPS: 0° 42' 1.6" S, 79° 59' 22.0" W) 7-26
with reconnaissance team members A. Caicedo, K. Rollins, C. Wood and other
local colleagues.

7.2.29 Chone Dam Middle: Active MASW array (GPS: 0° 41' 56.4" S, 79° 59' 23.1" W). 7-27

7.2.30 Chone Dam: Median theoretical Rayleigh wave dispersion curve; fit to. 7-30
experimental dispersion data (left). Median of top 1,000 Vs profiles from > 500,000
during inversion (right)

7.2.31 Chone Dam: Mean horizontal-to-vertical spectral ratio (HVSR). Theoretical lines 7-31
representing the H/V ratio curve and peak for median of top 1,000 profiles (next
version of this report).

7.2.32 Siltstone cores from boring P-1 at the APO1 accelerometer station. 7-29

7.2.33 APO1 station MAM and MASW Vs (GPS: 1° 2' 16.8" S, 80° 27' 36.0" W). 7-29

7.2.34 MAM/MASW at stations AMNT, ACHN, APED, AMA1 (clockwise from 7-30
top left: 0°56'27.6"S, 80°44'6"W; 0°41'52.8"S, 80°5'2.4"W; 0°4'4.8"S,
80°3'25.2"W; 0°56'6"S, 79°43'20"W).

7.2.35 MAM/MASW at AV18, ALIB, AMIL, AGYE (clockwise from top left: 7-31
0°18'46.8"S, 79°28'40.8"W; 2°14'34.8"S, 80°50'45.6"W; 2°10'51.6"S,
79°31'44.4"W; 2°3'14.4"S, 79°57'7.2"W).

7.2.36 AGYE1 and Catolica University stations MAM and MASW Vs measurements 7-32
(2°15'0.0"S, 79°54'36.0"W; 2°10'51.6"S, 79°54'14.4"W).

7.2.37 AGYE2 station (GPS: 2°11'56.4"S, 79°53'52.8"W). s and soil profile based on 7-33
previous subsurface investigations in 2005 and 2014 (Vera-Grunauer, 2014).

7.2.38 Manta AMNT station before the earthquake with locations of active and passive 7-34
surface wave arrays, HVSR, and AMNT. Base image: Digital Globe (©2016).

7.2.39 Manta AMNT station active MASW array (GPS: 0° 56' 27.6"S, 80° 44' 4.9"W). 7-34

7.2.40 Manta AMNT station: Median Rayleigh wave dispersion curve; fit to experimental 7-35
dispersion data (left). Median of top 1,000 Vs profiles from > 500,000 models in
inversion search (right).

7.2.41 Manta AMNT station: Mean horizontal-to-vertical spectral ratio (HVSR) obtained. 7-35
The red lines are the theoretical H/V spectral ratio curve and peak for the median of
the top 1,000 profiles.

7.2.42 Central Manta aero photo with locations of the seven HVSR test sites, HV1 to HV7. 7-36

7.2.43 Central Manta HVSR plots for sites HV1 to HV7. 7-37

7.2.44 Central Manta aero photo with locations of the HVSR tests and peaks for each site. 7-37

7.2.45 IESS Hospital before the earthquake with locations of active and passive surface wave 7-38
arrays and HVSR. Base image: Digital Globe (©2016).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xiii
7.2.46 IESS Hospital active MASW array (GPS: 0° 57' 15.5" S, 80° 43' 25.8" W). 7-38

7.2.47 IESS Hospital: Median Rayleigh wave dispersion curve; fit to experimental dispersion 7-39
data (left). Median of top 1,000 Vs profiles from > 500,000 models in inversion search
(right).

7.2.48 IESS Hospital: Mean horizontal-to-vertical spectral ratio (HVSR) obtained. The red 7-39
lines are the theoretical H/V spectral ratio curve and peak for the median of the top
1,000 profiles.

7.3.1 Strong Motion Stations (SMS) presented as color-coded circles according to the 7-41
recorded geo-mean PGA in the main event, with EW acceleration time histories and
major cities of Ecuador.

7.3.2 Earthquake damage assessment of main cities after the main event using ATC-20 7-42
by the Ministry of Housing. Colors indicate minor (green); partial (yellow); and major
(red) damage.

7.3.3 Average structural damage estimate expressed as a percentage among the 4 main 7-42
affected cities of Portoviejo, Pedernales, Manta, and Bahia de Caráquez, as function of
the number of floors. Data based on the Ministry of Housing post-earthquake assessment.

7.3.4 Manta AMNT station before the earthquake with locations of active and passive surface 7-43
wave arrays and HVSR (GPS: 0° 56' 27.6"S, 80° 44' 4.9"W). Base image: Digital
Globe (©2016).

7.3.5 Manta AMNT station: Comparison of active-source MASW and passive-source 2D 7-44
L-array MAM dispersion data (HRFK and F-K processing), used as raw input for the
Dinver inversion.

7.3.6 Manta AMNT station: Theoretical Rayleigh wave dispersion curves for top 1,000 7-44
velocity models from inversion. Theoretical dispersion curves with lowest misfit profile
from the 1,000 profiles.

7.3.7 Manta AMNT station: Median Vs profile based on the top 1,000 profiles (lowest misfit 7-45
of ~0.413) from over 500,000 models during inversion.

7.3.8 Manta AMNT station: Mean horizontal-to-vertical spectral ratio (HVSR) obtained. 7-46
The red lines are the theoretical H/V spectral ratio curve and peak for the median of the
top 1,000 profiles.

7.3.9 Manta AMNT station: Fourier amplitude spectrum of the recorded ground motion 7-47
during the main earthquake event.

7.3.10 Portovejo: Rapid assessment using ATC-20, with color tagging based on inspection 7-48
damage level (green: minor; yellow: partial; red: major). Map and data developed by
EC COE3.

7.3.11 Portoviejo: Average estimated damage in Portoviejo City (REFERENCE to be 7-49


provided in the next version of this report).

7.3.12 Portoviejo geomorphological map with ground zero as blue circle. Base map 7-50
developed by Instituto Espacial Ecuatoriano (IEE).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xiv
7.3.13 Soil profiles of SPT N60 and Vs, and rock cores at APO1 and Los Tamarindos sites. 7-51

7.3.14 Acceleration (Sa) and displacement spectra (Sd) spectra the record at APO1 station 7-51
during the main event and corresponding spectral amplification factor, Sa/PGA.

7.3.15 Geotechnical Zones of Guayaquil (Vera-Grunauer, 2014). 7-53

7.3.16 Typical subsoil profile at downtown Guayaquil. 7-55

7.3.17 Geotechnical description and Vs profile at station AGYE2. 7-56

7.3.18 Response spectra obtained for AGYE and AGYE2 records from 3 EC seismic events. 7-57

7.3.19 Response of ERU profile subjected to the 5 local ground motions of Table 7.3.2 in 7-58
terms of lateral displacement, max shear strain, ratios of G/Gmax, Vs /Vs,max, CSR,
Peak Ground Acceleration, PGA as functions of depths below the ground surface.

7.4.1 Flyover route along Portoviejo River were liquefaction was observed. 7-59

7.4.2 Liquefaction evidence along Portoviejo River. Photos from reconnaissance army 7-60
helicopter flyover (coordinates pending next version).

7.4.3 Mejia Bridge: Aerial view of embankment failure with assumed movement and failure 7-61
modes marked with yellow (after COE-3, GPS: 0°59'24"S, 80°28'11"W).

7.4.4 Boca de Briceno Bridge liquefaction evidence in free-field and adjacent to embankment 7-62
footing (top, est. GPS: 0°30'59"S, 80°26'31"W; bot. left, GPS: 0°30'59.1"S,
80°26'29.7"W, bot. right, GPS: 0°30'58.98"S, 80°26'28.93"W).

7.4.5 Manta overpass: Possible evidence of sand boils and cracking due to lateral movement 7-63
near the site (GPS: 0 o57'1.7"S, 80o43'8.7"W).

7.4.6 Signs of liquefaction in Tarqui, Manta (GPS: 0o57'9.1''S, 80o42'29.4''W). 7-63

7.4.7 Signs of liquefaction in Tarqui, Manta (GPS: left, 0o57'20.0''S, 80o42'39.9''W; 7-64
right, 0o57'20.4'', 80o42'38.4''W).

7.4.8 Liquefaction locations identified by COE-3 in Tarqui, Manta. 7-64

7.4.9 Liquefaction evidence from helicopter flyover of Manabí coast (GPS clockwise, top 7-65
left: 0o41'29.2''S, 80o15'10.7''W; 0o39'58.2''S, 80o18'17.2''W; 0o39'41.8''S,
80o18'46.6''W; 0o12'31.7''S, 80o15'22.5''W; 0o9'37.9''S, 80o15'47.2''W).

7.4.10 Aerial view of Manta Port with main geo-components (Digital Globe, ©2016). 7-66

7.4.11 Manta parking area: Aerial view of backfill behind rock berm (Digital Globe, ©2016). 7-67
Settlement and lateral displacement were measured by the reconnaissance team along
the orange lines.

7.4.12 Figure 7.4.12. Manta Port parking area: concentrated sand blow in the center of 7-68
pavement observed by GEER co-leader S. Nikolaou (GPS: 0°56'27.6"S, 80°43'29.4"W).

7.4.13 Manta Port parking area: Sand vents erupting from cracks in pavement formed by 7-68
lateral spreading. (left: 0°56'26.9"S, 80°43'29.5"W; right: 0°56'29.5"S, 80°43'27.8"W).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xv
7.4.14 Manta parking area: Vertical offset in ground retaining wall, wall deformation and light 7-69
pole rotation (left). Wide crack at the perpendicular (non-retaining) wall next to GEER
team member K. Rollins (right). Pole coordinates from GPS (0°56'26.7"S, 80°43'29.8"W).

7.4.15 Lateral movement of 40 cm at the 4-m tall retaining walls adjacent to boat ramp (top) 7-70
and heave and longitudinal cracking of base slab (bottom, GPS: 0°56'28.9"S,
80°43'26.8"W).

7.4.16 Manta Port parking area: Cumulative lateral spread displacements along the 5 lines 7-71
perpendicular to the shoreline shown on Fig. 4.7.11.

7.4.17 Manta Port parking area: Vertical offset in concrete pavement produced by lateral 7-71
spreading along the cyan Line A-A of Fig. 4.7.11, at 5-m intervals. Line A-A
is roughly parallel to the concrete wall at a distance of about 3.5 m from the wall face.

7.4.18 Manta Port wharfs: Reconnaissance team on motor boats (GPS 0°56'16.6"S, 7-72
80°43'29.3"W).

7.4.19 Manta Port wharf: (a) Rows of piles supporting the two piers on the northern end of 7-73
the port and (b) shear cracks in beam connecting a pile to the pile cap (GPS:
0°55'58.2"S, 80°43'18.8"W).

7.4.20 Manta Port wharf: (a) cracking and (b) compression failure of piles at the northwest 7-73
corner of the southern pier (GPS: 0°55'58.2"S, 80°43'18.8"W).

7.4.21 Manta Port wharf: Battered piles at 3-m OC supporting seaward side of the wharf with 7-74
alternate sloped 1H:5V batters with rock riprap slope behind them (GPS: 0°56'22.7"S,
80°43'33.8"W).

7.4.22 Manta Port wharf: Aerial view of pile-supported wharf along shoreline and areas with 7-75
pile damage. Offsets measured along the yellow line (GPS: 0°56'12.4"S, 80°43'23.5"W).

7.4.23 Manta Port wharf: Measured vertical and horizontal offset as a function of distance 7-75
along the yellow line parallel to the shore in Fig. 7.4.22.

7.4.24 Manta Port wharf: Cracking and shearing of pile head connection for backward-batter 7-76
piles, with minimal damage to forward-batter piles (GPS: 0°56'14.42"S, 80°43'27.41"W).

7.4.25 Manta Port wharf: Failure of pile head in backward-batter piles, with 30-40 cm 7-76
horizontal offset. Forward-batter piles were not damaged (GPS: 0°56'13.6"S,
80°43'26.6"W).

7.4.26 Manta Port wharf: Distress and cracking of pile head connection for piles with both 7-76
forward and backward batter (GPS: 0°56'11.2"S, 80°43'24.8"W).

7.4.27 Manta Port Authority Operations Building: Location on aerial photo of the port 7-77
(GPS: 0°55'56.3"S, 80°43'20.3"W).

7.4.28 Manta Port Authority Operations Building: Repairs performed in the embankment near 7-77
the embankment near the building at the time of reconnaissance (GPS: 0°55'56.3"S,
80°43'20.3"W).

7.4.29 Manta Port Authority Operations Building: (a) horizontal and vertical displacements 7-78
of embankment relative to building; (b) horizontal movement between building and

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xvi
embankment; (c) vertical settlement of embankment; (d) damage at the pile head
(GPS: 0°55'55.2"S, 80°43'19.6"W).

7.5.1 Locations of 7 retaining walls and masonry-type structures observed in reconnaissance. 7-79

7.5.2 Avenida 4 de Noviembre Bridge: Damage to east abutment and wing wall 7-80
(GPS: (a) 0°57'10.1''S, 80°43'5.5''W, (b) 0°57'10.2''S, 80°43'5.6''W, (c) 0°57'10.4''S,
80°43'5.1''W, (d) 0°57'10.1''S, 80°43'5.2''W, (e) 0°57'9.8''S, 80°43'04.8''W,
(f) 0°57'9.7''S, 80°43'4.6''W).

7.5.3 Avenida 4 de Nov. Bridge repairs on west abutment and wing wall: cracked concrete 7-81
is removed, exposing reinforcing steel (left: 0°57'15.2''S, 80°43'2.1''W, right:
0°57'15.2''S, 80°43'2.2''W).

7.5.4 Rio Chico (Las Chacras) Bridge: East abutments and wing walls (left, 7-82
0⁰58'35.6''S, 80⁰25'21.25''W); Displacement of toe slope at east abutment (right
0⁰58'34.96''S, 80⁰25'21.4''W).

7.5.5 Rio Chico (Las Chacras) Bridge: joint damage between approach and superstructure 7-82
(GPS: 0°58'35.6''S, 80°25'21.3''W). Team member Daniel Alzamora on the right.

7.5.6 Mejia Bridge: Design drawing from improvements at the Portoviejo-Crucita Highway 7-83
from the Ecuador Ministerio de Transporte y Obras Publicas (EMTOP, 2011).

7.5.7 Mejia Bridge: West abutment, gabions, wing wall without damage (top, 0°59'24.4''S, 7-84
80°28'11.5''W); Damage to east abutment, wing wall, and gabions (bot., 0°59'23.3''S,
80°28'11.9''W).

7.5.8 Mejia Bridge: Lateral displacements at east abutments (left: 0°59'23.7''S, 80°28'10.3''W, 7-85
right: 0°59'26.1''S, 80°28'09.7''W).

7.5.9 Boca de Briceno Bridge aerial photo (GPS: 0°31'3.7''S, 80°26'43.8''W). 7-85

7.5.10 Boca de Briceno Bridge: (a; b) Abutment and wing wall (GPS 0°30'58.4''S, 7-86
80°26'29.7''W; 0°30'57.7''S, 80°26'28.9''W); (c) Approach embankment and bridge
pavement looking south (GPS 0°30'58.1''S, 80°26'29.2''W). No visible damage observed.

7.5.11 Boca de Briceno Bridge: Damage to bridge joints and sidewalk (GPS: (a) 0°30'57.0''S, 7-86
80°26'29.7''W, (b) 0°30'58.5''S, 80°26'29.6''W, (c) 0°30'57.3''S, 80°26'30.2''W).

7.5.12 Bahia Bridge: Abutment and wing walls showing only minor damage to one of the 7-87
abutment walls (left, 0°36'37.2''S, 80°25'29.4''; right 0°36'37.7''S, 80°25'29.0''W).

7.5.13 Bahia Bridge: Bolted connections of wing walls (left 0°36'37.5''S, 80°25'29.7''W, 7-87
right 0°36'37.5''S, 80°25'29.7''W).

7.5.14 Bahia Bridge: Damage at intersection of abutment and north wing wall (left, 7-88
0°36'37.2''S, 80°25'29.5''W; right, 0°36'37.3''S, 80°25'29.5''W).

7.5.15 Bahía wall along Avenida Virgilio Ratti: Collapse of unreinforced stone masonry 7-89
(GPS: 0°36'29.8''S, 80°25'24.9''W).

7.5.16 Bahía wall along Avenida Virgilio Ratti: Wall deformation (left, 0°36'30.9''S, 7-89
80°25'25.2''W with GEER members R. Luque and G. Lyvers); Rotation and stacked

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xvii
walls (right, 0°36'30.3''S, 80°25'25.0''W with GEER members K. Rollins and C. Wood).

7.5.17 Bahía wall at Ave. Virgilio Ratti: Vertical deflection (left, 0°36'33.1''S, 80°25'25.6''W); 7-90
Wall bottom (center, 0°36'30.6''S, 80°25'25.0''W); Wall top (0°36'31.0''S, 80⁰25'25.2''W).

7.5.18 Aerial view of the collapsed wall at the Manta Port (GPS: 0°56'24.6''S, 80°43'22.0''W). 7-90

7.5.19 Manta Port wall: Damage and partial collapse (GPS: 0°56'28.5''S, 80°43'26.0''W). 7-91

7.5.20 Manta Port wall: Buckling of concrete boat ramp and collapse of wall corner 7-91
(GPS: 0°56'29.1''S, 80°43'26.7''W).

7.5.21 Manta Port wall: Lateral displacement of 40 cm (GPS: 0°56'30.6''S, 80°43'28.0''W). 7-91

7.5.22 Velboni supermarket wall in Portoviejo: Before (top, by Google Earth©) and after 7-92
the earthquake (bottom, by GEER team; GPS: 1°3'36.6''S, 80°27'29.6''W).

7.5.23 Wall at Portoviejo Velboni supermarket: Slope failure showing two primary slip 7-93
surfaces and wall rotating, following the slope movement (GPS: 1°3'37.8''S,
80°27'29.3''W). Team members A. Caicedo (foreground) and A. Zekkos, K. Rollis,
C. Wood, and an Ecuadorian colleague (background).

7.6.1 Landslides or rock falls reported by COE-3 along inspected roads. 7-94

7.6.2 Route followed during reconnaissance in blue line (Google Maps, ©2016). 7-95

7.6.3 Aerial photo of landslide on San Vicente-Canoa local road (GPS: 0°33' 5"S, 7-95
80°25'42"W).

7.6.4 Landslide near Canoa. Photo by COE-3, dated 4/20/16 (0°30'15"S, 80°26'44"W). 7-96

7.6.5 Landslide near Canoa. (COE-3 photo, GPS: 0°30'15"S, 80°26'4"W). 7-96

7.6.6 Landslide partially blocking the road near Jama (COE-3, GPS: 0°25'44"S, 80°26'59"W). 7-97

7.6.7 Percentile of relative impact of earthquake-induced (“co-seismic”) landslides for a total 7-98
of 540 landslides observed in Portoviejo, Bahía de Caráquez, Chone, Muisne and Crucita
(BGS, 2016).

7.6.8 Location of Loor and Navas macro-landslides in geologic map (Baldock, 1982). 7-98

7.6.9 Macro-landslides at Loor site (Ripalda, 2016) Base image: Google Earth ©2016. 7-99

7.6.10 Events of 4/16/16: Mw5.7 foreshock (left) and Mw7.8 main (right) (IG-EPN, 2016). 7-100

7.6.11 Loor site: Ground water observed at slope base (GPS: 0°24'55.2"S, 80°14'50.3"W). 7-100

7.6.12 Sequence of landslides at Loor site (Ripalda, 2016). 7-100

7.6.13 Modelling methodology to create a 3D model from images taken by drones during 7-101
reconnaissance developed by the GEER-ATC (developed by WSP|PB, Virtual Design
& Construction.

7.6.14 Loor site prior to the earthquake used to develop the 3D model (Google Earth, ©2016). 7-102

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xviii
7.6.15 Drone images of the Loor site landslides taken during reconnaissance after the 7-102
earthquake, used to develop the 3D model (Google Earth, ©2016).

7.6.16 3D model of Loor site developed by the GEER-ATC reconnaissance team digital 7-103
photos and drone images (WSP|PB, Virtual Design & Construction).

7.6.17 Section perpendicular hill slope at Loor site derived from the 3D model of 7-104
Fig. 7.6.14. Approximate pre-earthquake slope included in dashed yellow line.

7.6.18 Section along hill slope at Loor site derived from the 3D model of Fig. 7.6.14. 7-104
Approximate pre-earthquake slope included in dashed yellow line.

7.6.19 Loor site landslides: House at eastern end of the site destroyed in sunken area; 7-105
secondary (cane) house in lifted zone. Residual of white escarpment shale (GPS:
0°25'0.6"S, 80°15'40.1"W).

7.6.20 Loor site landslides: Destroyed ranch house and ground fissures around it (GPS: 7-105
0°25'0.1"S, 80°15'39.5"W).

7.6.21 House sinking along with landslide at Loor site (GPS: 0°25'0.2"S, 80°15'39.1"W). 7-106

7.6.22 Loor site landslides: Evidence of ground uplift (GPS: 0°25'0.18"S, 80°15'39.53"W). 7-106

7.6.23 Escarpment in the southwest side of Loor landslide (GPS: 0°25'2.3"S, 80°15'41.2"W). 7-107

7.6.24 Loor stables: Cracks (top, 0°25'1.4"S, 80°15'42.6"W; debris (bot., 0°25'0.8"S, 7-107
80°15'42.5"W).

7.6.25 Evidence of ground uplifting in Loor landslide (GPS: 0°24'51.24"S, 80°15'50.34"W). 7-108

7.6.26 Debris on the east side of the Loor landslide (GPS: 0°25'0.6"S, 80°15'37.9"W). 7-108

7.6.27 North view of the extreme east of the landslide (GPS: 0°25'0.1"S, 80°15'38.2"W). 7-108

7.6.28 Navas site macro-landslides (Ripalda, 2016). Base image: Google Earth (©2016). 7-109

7.6.29 Navas landslide west site with Isla Central at right. (GPS: 0°24'41.0"S, 80°13'28.1"W). 7-109

7.6.30 Photo at the eastern limit of the Navas landslide (GPS: 0°24'41.0"S, 80°13'28.0"W). 7-109

7.7.1 Ecuador Transportation Network (Academic, 2016). 7-110

7.7.2 Global Competitiveness Report (2015) Infrastructure: Quality of Road Ratings. 7-111

7.7.3 Concrete pavement damage south of the town of Pedernales (GPS: 7-111
0°2'5.4204"S, 80°4'27.3216"W).

7.7.4 Briceno Bridge location. The original U-shape old roadway alignment was 7-112
replaced by this bridge and new alignment. (Google Earth Image ©2016 TerraMetrics).

7.7.5 Boca de Briceno Bridge aerial photo (GPS: 0⁰ 30'58.27''S, 80⁰ 26'29.08''W). 7-112

7.7.6 Boca de Briceno Bridge design layout (EMTOP, 2010b). 7-113

7.7.7 Boca de Briceno embankment geometry (GPS: 0°30'56.95"S, 80°26'30.08"W). 7-114

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xix
7.7.8 Boca de Briceno Bridge geotechnical design parameters (Asesorias y Estudio 7-114
Tecnico C. LTDA, Undated).

7.7.9 Construction sequence using Geopiers (Asesorias y Estudio Tecnico 7-115


C. LTDA, Undated).

7.7.10 Boca de Briceno Bridge location (GPS 0°30'57.0"S, 80°26'30.1"W) with respect to 7-116
strong ground motion records the main earthquake epicenter. Base Image: Google
Earth (©2016).

7.7.11 Boca de Briceno Bridge: Geopier design layout by Station (Asesorias y Estudio Tecnico, 7-117
undated). Note English translation Espaciamiento (Spacing) and Profundidad (Depth).

7.7.12 Geopier design layout: Top - Typical section; Bot. - as built plans (EMTOP, 2014). 7-117

7.7.13 Boca de Briceno Bridge construction: Left - equipment installing geopiers; right - top 7-118
of a completed 60-cm dia. geopier (est. GPS: 0°30'59"S, 80°26'29"W). Photos courtesy
of Pivaltec.

7.7.14 Liquefaction and its effects next to Boca de Briceno Bridge embankment (top: est. 7-118
GPS: 0°30'6"S, 80°26'3"W; bot. left: 0°30'59.1"S, 80°26'29.7"W; bot. right:
0°30'59.0"S, 80°26'28.9"W).

7.7.15 Boca de Briceno Bridge: minor pavement damage (est. GPS: top 0°31'17"S, 7-119
80°26'24"W; bot: 0°31'12"S, 80°26'25"W).

7.7.16 Boca de Briceno Bridge: overall good performance of the embankment with little to no 7-120
vertical displacement (GPS: 0°30'57.0"S, 80°26'30.1"W).

7.7.17 Manta Overpass: Cracking and ground movement (GPS: 0o57'1.2''S, 80o43'13.2''W). 7-120

7.7.18 Cracked sidewalk at the Manta Overpass site (GPS: 0o57’01.16”S, 80o43’13.24”W). 7-121

7.7.19 Possible evidence of sand boils and cracking due to lateral movement near the Manta 7-121
Overpass site (GPS: 0o57'01.7''S, 80o43'08.7''W).

7.7.20 Cracking and ground movement at and around the Mobil gas station near the Manta 7-122
Overpass site (GPS: 0o56'59.1''S, 80o43'14.8''W).

7.7.21 Cracking and ground settlement around the underground tanks at the Mobil gas station 7-122
near the Manta Overpass site (GPS: 0 o56'59.1''S, 80o43'14.8''W).

7.7.22 Cracking, ground movement north of Mobil station (GPS: 0o56'53.6''S, 80o43'16.0''W). 7-122

7.7.23 Location of Puente Mejia Bridge (Google Maps ©2016). 7-123

7.7.24 Mejia Bridge prior to the earthquake (Google Maps, ©2016). 7-123

7.7.25 Plan of the Mejia Bridges (EMTOP, 2011). 7-124

7.7.26 Typical longitudinal section of the Mejia Bridges (EMTOP, 2011). 7-124

7.7.27 Steel Beam and Concrete Deck Detail (EMTOP, 2011). 7-124

7.7.28 Subsurface investigations at Mejia Bridge (Google Maps, ©2016). 7-125

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xx
7.7.29 Pre-earthquake Boring B-1 at Mejia Bridge site. 7-126

7.7.30 Pre-earthquake Boring B-2 at Mejia Bridge site. 7-127

7.7.31 Post-earthquake Boring P-01 at Mejia Bridge site. 7-128

7.7.32 Mejia Bridge site: Post-earthquake CPTu results. 7-129

7.7.33 Mejia Bridge site: Post-earthquake Vs measurements from CPTu. 7-129

7.7.34 Mejia Bridge site: Post-earthquake Vs measurements from ReMi+MASW. 7-130

7.7.35 Mejia Bridge approach embankment: Aerial views of failure by COE-3 (GPS: top, 7-131
0°59'24"S, 80°28'11"W; bottom, 0°59'24"S, 80°28'11"W).

7.7.36 Mejia Bridge: Road displacement (GPS: 0°59'24"S, 80°28'11"W, Geoestudios). 7-132

7.7.37 Bahia-San Vicente, Los Caras Bridge. Photo by Ecuador Army Corps of Engineers 7-132
looking eastward from San Vicente to Bahía de Caráquez (GPS: 0°36'30.7"S,
80°24'30.6"W).

7.7.38 Aerial view of Los Caras (Bahía) Bridge by Ecuador Army Corps of Engineers 7-133
(GPS: 0°36'33.6"S, 80°24'58.7"W).

7.7.39 Los Caras (Bahía) Bridge cross-section (Ecuador Army Corps of Engineers, 2010). 7-134

7.7.40 Los Caras (Bahía) Bridge soil profile (Ecuador Army Corps of Engineers, 2010). 7-134

7.7.41 Ecuador Army Corps of Engineers team members show the MIDUVI inspection 7-135
document of the Los Caras (Bahía) Bridge (GPS: 0°36'48.9"S, 80°24'46.9"W).

7.7.42 Los Caras (Bahía) Bridge: Modular expansion joint, Bahía de Caráquez side 7-136
(GPS: 0°36'48.85"S, 80°24'46.93"W), by Romo (2016).

7.7.43 Los Caras (Bahía) Bridge: Modular expansion joint, San Vincente side (GPS: 7-136
0°36'30.74"S, 80°24'30.54"W), by Romo (2016).

7.7.44 Los Caras (Bahía) Bridge: Neoprene rubber standard expansion joint (GPS: 7-136
0°36'35.28"S, 80°25'17.15"W), by Romo (2016).

7.7.45 Los Caras (Bahía) Bridge: External movements synchronizers (GPS: 0°36'33.6"S, 7-137
80°24'58.7"W), by Romo (2016).

7.7.46 Los Caras (Bahía) Bridge: Cracking of Pier 3 on the horizontal shear block (Bahia side). 7-137
The pier has elastomeric bearings (GPS: 0°36'36.6"S, 80°25'27.3"W), by EC Army
Corps of Engineers.

7.7.47 Friction pendulum device as part of the base isolation system of Los Caras Bridge 7-138
(Left), Inner displacement isolator view (GPS: 0°36'33.58"S, 80°24'58.68"W), by Romo,
2016.

7.8.1 Locations of four inspected dams as yellow pins. Base image: Google Earth (©2016). 7-139

7.8.2 La Esperanza Dam aerial view from helicopter flyover (GPS: 0°53' 2.5''S, 80°5'3.0''W). 7-140

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxi
7.8.3 La Esperanza Dam: Embankment cross section (Rehabilitation Center, Manabí, 1983). 7-141

7.8.4 La Esperanza Dam: Plan view (Rehabilitation Center, Manabí, 1983). 7-141

7.8.5 La Esperanza Dam: Abutment drainage galleries (Rehabilitation Center, Manabí, 1983). 7-142

7.8.6 La Esperanza Dam: Intake tower and conduit drawings (Rehab. Center, Manabí, 1983). 7-142

7.8.7 Spillway section at right abutment (Rehabilitation Center, Manabí, 1983). 7-142

7.8.8 La Esperanza Dam: Right and left drainage galleries after the main seismic event. 7-143
Photos from the MIDUVI Inspection Commission (GPS: 0° 53' 19.0'' S, 80° 4' 35.3'' W).

7.8.9 La Esperanza Dam: Differential movement of intake tower and access bridge. Photo 7-144
from the MIDUVI Inspection Commission (GPS: 0° 53' 14.3'' S, 80° 4' 30.4' W).

7.8.10 La Esperanza Dam: Damage to columns supporting the intake tower gate hoist. Photo 7-144
from the MIDUVI Inspection Commission (GPS: 0° 53' 14.5'' S, 80° 4' 30.0'' W).

7.8.11 La Esperanza Dam: Spillway radial gates. Photo from the MIDUVI Inspection 7-145
Commission (GPS: 0° 53' 9.2'' S, 80° 4' 23.6'' W).

7.8.12 La Esperanza Dam: A spillway radial gate. Photo from the MIDUVI Inspection 7-145
Commission (GPS: 0°53' 8.9'' S, 80° 4' 36.4'' W).

7.8.13 Poza Honda Dam aerial view. Base Image: Google Earth (©2016). 7-146

7.8.14 Poza Honda Dam. MIDUVI Inspection photo (GPS: 1 °6' 39.8'' S, 80° 12' 32.5'' W). 7-146

7.8.15 Poza Honda Dam: Transverse cracks at crest. MIDUVI Inspection Commission photo 7-147
(GPS: 1° 06' 51.7'' S, 80° 12' 16.5'' W).

7.8.16 Poza Honda Dam: Cracks in upstream concrete slabs. MIDUVI Inspection Commission 7-148
photo (GPS: 1° 6' 51.8'' S, 80° 12' 15.8'' W).

7.8.17 Poza Honda Dam: Drainage ditch at toe with no seepage at the time of inspection. 7-148
MIDUVI Inspection Commission photo (GPS: 1° 6' 51.1'' S, 80° 12' 17.1'' W).

7.8.18 Poza Honda Dam: No seepage observed in the relief wells at the time of inspection, 7-149
MIDUVI Inspection Commission photo (GPS: 1° 6' 46.7'' S, 80° 12' 17.8'' W).

7.8.19 Poza Honda Dam: Uncontrolled spillway and stilling basin. MIDUVI Inspection 7-149
Commission photo (GPS: 1° 6' 52.2'' S, 80° 12' 17.0'' W).

7.8.20 Poza Honda Dam: Low flow outlet at time of inspection. MIDUVI Inspection 7-149
Commission photo (GPS: 1° 6' 46.2'' S, 80° 12' 17.8'' W).

7.8.21 Chone (Rio Grande) Dam aerial view from flyover (GPS: 0°41'44.1''S, 79°59'30.6''W). 7-150

7.8.22 Geologic Profile of Chone Dam location. (Chávez, 2013). 7-151

7.8.23 Chone Dam: Plan view after Chávez (2013). 7-151

7.8.24 Chone Dam: Typical dam cross section after Chávez (2013). 7-152

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxii
7.8.25 Chone Dam: Grout curtain and cutoff wall after Chávez (2013). 7-152

7.8.26 Instrumentation at Chone Dam (ESPE-Innovativa & Empresa Publica del Agua, 2015). 7-153

7.8.27 Chone Dam: Wet spot at the toe (GPS: 0° 41' 43.2'' S, 79° 59' 29.6'' W). 7-154

7.8.28 Chone Dam: Concrete spillway (GPS: 0° 41' 50.5'' S, 79° 59' 36.6'' W). 7-154

7.8.29 Chone Dam: Spillway cracks in slab, wall, counterfort (GPS 0°42'2.0''S, 7-155
79°59' 25.9''W).

7.8.30 Chone Dam: Spillway crack in wall and counterfort (GPS: 0° 42' 2.1''S, 7-155
79° 59' 25.6''W).

7.8.31 Chone Dam: Gate seal at low flow outlet (MIDUVI, GPS: 0° 42' 5.6'' S, 7-156
79° 59' 21.5''W).

7.8.32 Chone Dam: Slope instability at right abutment (GPS: 0° 41' 40.7'' S, 79° 59' 17.7' 'W). 7-156

7.8.33 Slope instability at right abutment above dam (GPS: 0° 42' 1.6'' S, 79° 59' 10.9'' W). 7-157

7.8.34 Slope instability, right abutment, above dam. MIDUVI Inspection Commission photo 7-157
(GPS: 0° 41' 59.0'' S, 79° 59' 21.1'' W).

7.8.35 Slope instability, left abutment by concrete spillway. MIDUVI Inspection Commission 7-157
photo (GPS: 0°42'5.0'' S, 79°59'25.6'' W).

7.8.36 San Vicente Dam: Aerial view. Base Image: Google Earth ( ©2016). 7-158

7.8.37 San Vicente Dam: Uncontrolled spillway. Base Image: Google Earth ( ©2016). 7-158

7.8.38 San Vicente Dam: Uncontrolled spillway crest. MIDUVI Inspection Commission photo 7-159
(GPS: 2° 0' 1.8'' S, 80° 31' 54.8'' W).

7.8.39 San Vicente Dam: Spillway chute slabs (MIDUVI, GPS: 2° 0'’ 1.6'' S, 80° 31' 56.1'' W). 7-160

7.8.40 San Vicente Dam: Repaired spillway joint (MIDUVI, GPS: 2°0'1.6''S, 80°31' 56.1''W). 7-160

7.8.41 San Vicente Dam: Shear crack at spillway wall joint, (MIDUVI, GPS: 7-160
2°0' 1.4''S, 80°31'54.6''W).

7.8.42 San Vicente Dam: Shear crack at spillway training wall joint. Top of MIDUVI 7-161
Inspection Commission photo (GPS: 2° 0' 1.3'' S, 80° 31' 55.8'' W).

7.8.43 San Vicente Bridge, Santa Elena. MIDUVI Inspection Commission photo (GPS: 7-161
2° 0' 1.8'' S, 80° 31' 54.4'' W).

7.8.44 San Vicente Dam: Spillway bridge joint separation of 20 cm at left support. 7-162
MIDUVI Inspection Commission photo (GPS: 2° 0' 2.3'' S, 80° 31' 54.9'' W).

7.8.45 San Vicente Dam: Spillway bridge joint separation of 7 cm at right support. 7-162
MIDUVI Inspection Commission photo (GPS: 2° 0' 1.36'' S, 80° 31' 54.7'' W).

7.9.1 Soil classification from two borings at Rio Chico (Las Chacras) Bridge embankments 7-164
(after LUP Geotechnical Report, 2013).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxiii
7.9.2 Undrained shear strength Su from two borings at the Mejia Bridge (NYLIC, 2001). 7-166

7.9.3 Embankment fill: Rio Chico (top left, GPS: 0°58'34.5"S, 80°25'20.0"W); Mejia (top 7-167
right, GPS: 0°59'25.4"S, 80°28'10.2"W) and its geotextile, geogrid (bot., GPS:
0°59'25.8"S, 80°28'10.5"W).

7.9.4 Failure modes: circular (left) and sliding block (right) after Samtani & Nowatzki (2006). 7-168

7.9.5 Rio Chico embankment settlement (GPS: 0°58'34.78"S, 80°25'20.21"W). 7-168

7.9.6 Rio Chico: Settlement and rotation of block in foreground and circular global failure 7-169
in background (GPS: 0° 58' 36.6" S, 80° 25' 23.7" W).

7.9.7 Rio Chico: Settlement and rotation of block in background and circular global failure 7-169
in foreground (GPS: 0°58'38.09"S, 80°25'25.67"W).

7.9.8 Rio Chico embankment: Block settlement and rotation (GPS: 0°58'41.2"S, 7-170
80°25'26.8"W). GEER member K. Rollins shown on background.

7.9.9 Rio Chico embankment: Separation of pipe (GPS: 0°58'43.5"S, 80°25'26.7"W). 7-170

7.9.10 Crack depth measurements along the Rio Chico embankment of 3.7 m (left, GPS: 7-171
0°58'72.0''S, 80°25'26.6''W) and 2.1 m (right, GPS: 0°58'34.9''S, 80°25'20.7''W).

7.9.11 Mejia Bridge and approaches plan view during construction. Arrows show the direction 7-171
of movement. Base image from Google Earth (©2016 DigitalGlobe).

7.9.12 Mejia Bridge: Aerial view of embankment failure with assumed movement and failure 7-172
modes marked with yellow (after COE-3, GPS: 0°59'24"S, 80°28'11"W).

7.9.13 Mejia Bridge embankment: View after earthquake (top, modified from 7-172
Weiser-Woodward, 2016); Repairs to approach to allow access (bot., GPS:
0°59'26.24"S, 80°28'10.19"W).

7.9.14 Mejia Bridge: Failures of SW abutment and gabions (GPS: 0°59'23.5"S, 80°28'11.6"W). 7-173

7.9.15 Mejia Bridge: North abutment performance (GPS: 0°59'24.94"S, 80°28'11.26"W). 7-173

7.10.1 IESS Hospital post-earthquake. GEER-ATC team visit on 4/28/16. Team member 7-174
G. Lyvers and X. Vera on the right (GPS: 0°57'18.5"S, 80°43'26.5"W).

7.10.2 IESS Hospital location (GPS: 0°57'17.8"S, 80°43'25.8"W). Aerial photo 7-175
pre-earthquake (Digital Globe, ©2016).

7.10.3 IESS Hospital: Locations of active and passive surface wave arrays, and HVSR in-situ 7-175
testing performed by GEER on 4/28/16. Base image: Digital Globe (©2016).

7.10.4 IESS Hospital: Active MASW array (GPS: 0°57'15.5"S, 80°43'25.8"W). 7-176

7.10.5 IESS Hospital: Passive MAM array (GPS: 0°57'16.0"S, 80°43'27.6"W). 7-176

7.10.6 IESS Hospital: HVSR testing (GPS: 0°57'15.9"S, 80°43'25.8"W). 7-177

7.10.7 IESS Hospital: Median theoretical Rayleigh wave dispersion curve for the top 1,000 7-178
profiles; fit to experimental dispersion data (left). Median of the top 1,000 V s profiles

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxiv
obtained from over 500,000 models searched during inversion (right).

7.10.8 IESS Hospital: Mean horizontal-to-vertical spectral ratio (HVSR). Red theoretical 7-178
lines represent the HVSR curve and peak for the median of the top 1,000 profiles.

8.1.1 Residential buildings under construction with 2-way waffle slabs a) Building 8-1
in Bahía de Caraquez. (GPS coordinates: 0°37’30.7”S, 80°25’31.5”W).
b) Building in Manta (GPS coordinates: 0°57’8.1”S, 80°42’26.9”W).

8.1.2 a) Residential building under construction with 2-way waffle slabs in Tarqui, 8-2
Manta. (GPS coordinates: 0°57'28.67"S, 80°42'32.41"W). b) Details of
reinforcement layout in the back of the house.

8.1.3 a) Residential building under construction with 1-way beams and waffle slabs 8-2
in Tarqui, Manta. (GPS coordinates: 0°57’27.3”S, 80°42’32.9”W).

8.1.4 Rodriguez Zambrano Hospital building with moment resisting concrete frames 8-3
and waffle slabs in Manta. (GPS coordinates: 0°57’21.8”S, 80°44’30.8”W).

8.1.5 IESS Hospital building with moment resisting concrete frames and flat slabs in 8-3
Manta. (GPS coordinates: 0°57’18.0”S, 80°43’25.2”W).

8.1.6 The 911 ECU center in Portoviejo. It is a 3-story steel framed building (GPS 8-4
coordinates: 1°4’16.5”S, 80°26’47.5”W).

8.1.7 Wood framed house with metal deck roof in Canoa (GPS coordinates: 8-4
0°27’47.7”S, 80°27’16.9”W).

8.1.8 The lead rubber isolators under the Sky Building which is under construction in 8-5
Guayaquil. The reconnaissance team did not observe the structure. However,
in the building’s marketing website, it says it was visually inspected and no
damages were observed after the earthquake. (GPS coordinates:
2°8’39.7”S, 79°52’57.2”W - Photo obtained from http://skybuilding.com.ec/)

8.1.9 The Bahía to San Vicente Bridge is a seismic isolated bridge using triple friction 8-5
pendulum isolators. See section 8.10.1 for more information (GPS
coordinates: 0°36’35.9”S, 80°25’25.6”W).

8.1.10 The Guayas and Quil monument (86 tons), which will be completed in July of 8-6
2016 in Guayaquil, is base isolated. (GPS coordinates: 2°9’16.7”S,
79°52’42.6”W – Photo obtained from www.ecuavisa.com).

8.2.1 Building materials near Los Esteros Market in Tarqui, Manta. a) Fine aggregate 8-7
with sea shells. b) Good quality coarse aggregate (GPS coordinates: 0°57'8.94"S,
80°42'29.38"W).

8.2.2 Signs of severe corrosion on reinforcing steel previous to the earthquake. a) 8-8
Collapsed electric pole with signs of reinforcement corrosion and reduction of
effective area of longitudinal steel in Tarqui, Manta (GPS coordinates:
0°57'29.92"S, 80°42'36.45"W). b) Signs of severe corrosion in the steel at the base
of a 1st story column of a 3-story building tagged for demolition in Tarqui, Manta
(GPS coordinates: 0°57'9.04"S, 80°42'41.54"W).

8.2.3 New construction where the form blocks have been left in the waffle slab in 8-9

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxv
Tarqui, Manta (GPS coordinates: 0°57’27.3”S, 80°42’32.9”W).

8.2.4 a) A 1st floor column in the 3-story Hotel Las Gaviotas in Tarqui, Manta. 8-10
There is significant spalling associated to both, previously corroded steel and
the seismic demand imposed by the earthquake (GPS coordinates: 0°57’5.7”S,
80°42’48.5”W). b) Small-diameterhoop spacing details of a 1st floor unconfined
column in a 4-story commercial building in Tarqui, Manta (GPS coordinates:
0°57’14.9”S, 80°42’53.9”W).

8.2.5 Damage at the base of 6-story mixed-occupancy building in Tarqui, Manta. 8-11
There is significant spalling at the bottom of the first floor columns, showing
smooth longitudinal rebar and hoop spacing of approximately 2/3 of the column
width. (GPS coordinates: 0°57'14.55"S, 80°42'53”W).

8.2.6 a) Failure of first floor columns in building under construction in Canoa. The 8-12
column has large stirrup spacing. b) The slab to column connection at the edge
shows a lack of dowels, therefore providing no structural integrity (GPS
coordinates: 0°27'42.67"S, 80°27'20.67"W).

8.2.7 Slab reinforcement detail in a residential building under construction in Tarqui, 8-13
Manta. a) General view of joist, beams, and slab reinforcement; b) Details of
beam reinforcement near a beam-column joint (GPS coordinates:
0°57'28.67"S, 80°42'32.41"W).

8.2.8 Column with spalled-off concrete cover and evidence of healthy a core thanks 8-14
to closely spaced perimeter hoops and cross-ties. Zero Zone in Portoviejo.
(GPS coordinates: 1°3'13.53"S, 80°27'8.60"W)

8.2.9 Column with spalled-off concrete cover and evidence of healthy a core thanks 8-14

8.3.1 Seismic zonation of Ecuador in the CPE INEN CEC-2001. 8-17

8.3.2 Design spectra in the CEC-2001. 8-18

8.3.3 Seismic zonation for Ecuador in the 2015 Ecuador Building Code. 8-19

8.3.4 Design spectrum in the NEC-2015. 8-20

8.3.5 Displacement spectrum in the CEC-2015. 8-20

8.4.1 Green tag used to evaluate post-earthquake safety. It says that the structure has 8-22
been inspected and no structural damages or risks are apparent. Tags were
provided by MIDUVI.

8.4.2 Yellow tag used to evaluate post-earthquake safety. It says that the structure 8-23
has been inspected. The structural engineer would list all of the damages as well
as the areas that are restricted. Tags were provided by MIDUVI.

8.4.3 Red tag used to evaluate post-earthquake safety. It says that the structure has 8-23
been inspected and there is structural damage and risk. Entering the building
is restricted unless a person brings a legal document saying otherwise. It warns
that entering the property could cause injuries or even death.

8.4.4 Damage assessment form used by structural engineering professionals to 8-24

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxvi
record the structural damage of the property.

8.4.5 This shows the damage assessment in the affected cities until April 29, 2016. 8-25
The survey was completed in the zone 0 and in the most affected areas. The
damage survey and chart were provided by COE3, MIDUVI.

8.4.6 Map of colored tags of properties in Manta after post-earthquake safety 8-26
evaluation. The map was provided by COE3, MIDUVI.

8.4.7 Map of colored tags of properties in Manta after post-earthquake safety 8-27
evaluation. Information updated until April 29, 2016. The map was provided by
COE3, MIDUVI.

8.4.8 Probability of Failure. Comparison between US Code, Chile Maule 8-27


Earthquake and ECU M7.8 Ecuador Earthquake updated data until April 29th.
Data from ATC 63, Chris Poland ASCE31/41 webinar and COE3, MIDUVI
Ecuador.

8.5.1 Residential building in Tarqui, Manta, where the exterior walls collapsed. The 8-30
roof also fell in this property. (GPS coordinates: 0°57’14.6”S, 80°42’38.5”W).

8.5.2 a) Collapsed brick walls in building in IESS Hospital in Manta (GPS 8-31
coordinates: 0° 57' 18.0"S, 80° 43' 26.47"W). b) Concrete building with wood
and metal deck extension in Portoviejo. Non-structural wall completely collapsed
(GPS coordinates: 0°57’14.6”S, 80°42’38.5”W). c) Concrete building in Canoa
where brick and masonry exterior walls collapsed. (GPS coordinates:
0°27'46.59"S, 80°27'23.56"W).

8.5.3 This is a site of new construction. These are dowels at the concrete columns 8-31
to be attached to the non-structural walls (GPS coordinates: 0°57'7.31"S,
80°42'26.44"W).

8.5.4 Soft story failure: a) The first floor of this building in Manta collapsed. Some 8-32
of the rebar in the columns that failed was corroded. Additionally, the columns
were not confined (GPS coordinates: 0°57’14.8”S, 80°42’31.6”W). b) The first
floor of this building in Tarqui, Manta collapsed. Column confinement was
not apparent (GPS coordinates: 0°57’29.9”S, 80°42’42.1”W). c) The second floor
of this building in Portoviejo collapsed (GPS coordinates: 0°57’29.9”S,
80°42’42.1”W).

8.5.5 Plastic hinging in the 1st floor columns of this residential building with 1st 8-33
floor commercial occupancy in Tarqui, Manta (GPS coordinates: 0°57’7.1”S,
80°42’55.3”W).

8.5.6 Plastic hinges formed in the 1st floor columns of this building in Manta 8-33
(GPS coordinates: 0°57’9.5”S, 80°42’49.4”W).

8.5.7 a) The 4th and 5th floors collapsed in this residential building in Portoviejo 8-34
(GPS coordinates: 1°3'13.53"S, 80°27'8.60"W). b) The concrete blocks
restrained the column height, causing captive column failure for this column.

8.5.8 a) Rebar corrosion in a beam-column connection, and slab reinforcement in a 8-35


3-story building located in Tarqui, Manta (GPS coordinates: 0°57'7.94"S,
80°42'41.90"W). b) Rebar area reduction at the base of a column in a 2-story

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxvii
house located in Tarqui, Manta (GPS coordinates: 0°57'19.77"S, 80°42'40.68"W).

8.5.9 a) Pounding of independent buildings at the Bahia Airport (GPS 8-36


coordinates: 0°36’5.3”S, 80°24’27.0”W). b) Pounding of two buildings in
Portoviejo. The reconnaissance team observed that the roof metal deck is
continuous at the separation (GPS coordinates: 1°3’11.6”S, 80°27’8
c) Pounding evidence in Hotel Las Gaviotas in Tarqui, Manta (GPS
coordinates: 0°57'6.08"S, 80°42'48.85"W).

8.5.10 A house that settled more than a foot in Manta. a) Evidence of the settlement 8-37
with respect to an adjacent house. b) Interior of the house (GPS coordinates:
0°57'15.5"S, 80°42'33.7"W).

8.5.11 House settlements and soil failure evidence in Tarqui, Manta. (GPS 8-37
coordinates: 0°57'18.0"S, 80°42'38.9"W).

8.5.12 Sideway collapse of residential structure in Pedernales (GPS coordinates: 8-38


0°4'4.8"N, 80°3'23.5"W).

8.5.13 Building collapse in Pedernales. The columns shown have hinges at both ends 8-38
(GPS coordinates: 0°4'3.94"N, 80°3'24.6"W).

8.5.14 Collapse of building with flat slab construction in Pedernales (GPS 8-39
coordinates: 0°4'2.5"N, 80°3'33.9"W).

8.5.15 Collapse of the Unidad Educativa “Linus Pauling” building in Manta (GPS 8-39
coordinates: 0°57'24.4"S, 80°42'40.8"W). See section 8.8 for more information
regarding this structure.

8.5.16 Building that was under construction collapsed. The construction type was 8-40
reinforced concrete 2-way flat slabs supported on columns (GPS coordinates:
0°27'43.8"S, 80°27'21.0"W).

8.6.1 Concrete stair damaged and blocked by fallen infill walls at the IESS Hospital, 8-41
Manta. (GPS coordinates: 0°57'17.8''S, 80°43'25.4''W).

8.6.2 Concrete stairs damaged in Residential houses, Pedernales. (GPS coordinates: 8-42
0°4'12.9''N, 80°3'18.5''W).

8.6.3 Concrete stairs damaged, building in construction, Portoviejo. (GPS 8-42


coordinates: 0°59'48.42'S, 80°28'0.4''W).

8.6.4 Glazing damaged at the IESS Hospital, Manta. (GPS coordinates: 0°57'17.8''S, 8-43
80°43'25.4''W).

8.6.5 Glazing damaged at residential houses, Pedernales (GPS 0°4'14.5''N, 8-43


80°3'15.3''W).

8.6.6 Examples of commonly observed failure in exterior masonry infill walls. 8-44
Universidad Técnica de Manabi in Portoviejo (GPS coordinates:
1°2'39''S, 80°27'21''W).

8.6.7 Out of plane failure of exterior masonry infills at IESS Hospital, Manta 8-44
(GPS coordinates: 0°57'17.8''S, 80°43'25.4''W).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxviii
8.6.8 In plane failure of interior hollow clay masonry partitions that then led to 8-45
out-of-plane failures at IESS Hospital, Manta (GPS coordinates: 0°57'17.8''S,
80°43'25.4''W).

8.6.9 Cantilever wall damaged at “Fuerte Militar Manabi” Portoviejo (GPS 8-45
coordinates: 1°3'6.6''S, 80°28'30.8''W).

8.6.10 Partition damaged at Eloy Alfaro International Airport, Manta (GPS 8-46
coordinates: 0°57'11.34''S, 80°41'3.4''W). Method for mitigating damage to
partitions by installing angles connected to the slab (FEMA E-74).

8.6.11 Precast concrete lamp posts damaged at Tarqui, Manta (GPS coordinates: 8-47
0°57'27.7''S, 80°42'35''W).

8.6.12 Recessed light fixtures and ceiling damaged at Eloy Alfaro International 8-47
Airport, Manta (GPS coordinates: 0°57'11.34''S, 80°41'3.4''W).

8.6.13 In plane failure of interior hollow clay masonry partitions that then led to 8-48
out-of-plane failures at IESS Hospital, Manta (GPS coordinates: 0°57'17.8''S,
80°43'25.4''W).

8.6.14 Mitigation option to provide diagonal bracing for ceiling systems in 8-48
FEMA E-74.

8.6.15 Parapet damaged at the Bahia de Caraquez “Los Perales” Airport, Bahia de 8-49
Caraquez. And truck damaged by the parapet piece. (GPS coordinates: 0°36'5.2''S,
80°24'29''W).

8.6.16 Parapet damaged at the Jaramijo Port Building, Jaramijo. (GPS 8-49
coordinates: 0°56'38.6''S, 80°38'16.1''W) and Unidad educative San Jose
(GPS coordinates: 0°57’14.9''S, 80°42'45''W).

8.6.17 Veneer damaged at the IESS Hospital, Manta. (GPS coordinates: 0°57'17.8''S, 8-50
80°43'25.4''W).

8.6.18 Pipes and ducts fallen from the ceiling at IESS Hospital, Manta. (GPS 8-51
coordinates: 0°57'17.9''S, 80°43'26.37''W).

8.6.19 HVAC equipment on the roof IESS Hospital, Manta. (GPS coordinates: 8-51
0°57'17.9''S, 80°43'26.37''W).

8.6.21 AC unit attachment at the residential area, Manta. (GPS coordinates: 8-52
0°57'29.3''S, 80°42'36.4''W).

8.6.22 Furniture displacement at the HRZ Hospital, Manta. April 17th. (GPS 8-52
coordinates: 0°57'15.3''S, 80°44'32.8''W).

8.6.23 File cabinets at the HRZ Hospital, Manta. April 17th. (GPS coordinates: 8-53
0°57'15.3''S, 80°44'32.8''W).

8.6.24 File cabinets at the HRZ Hospital, Manta. April 28th. Bracing to walls and 8-54
between cabinets. (GPS coordinates: 0°57'15.3''S, 80°44'32.8''W).

8.6.25 Light fixture anchorage in good condition and Gas tubes chained to walls at the 8-54

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxix
HRZ Hospital, Manta. April 17th. (GPS coordinates: 0°57'21.8''S, 80°44’308''W).

8.7.1 a) Typical residential building with a commercial ground floor in Manta 8-56
(GPS coordinates: 0°57’14.4”S, 80°42’36.6”W). b) Residential building in
Tarqui, Manta (GPS coordinates: 0°57’16.83”S, 80°42’32.1”W).

8.7.2 a) Typical residential building where floors were added to the original 8-57
structure. The third floor seems to have been added after the original construction
of the two story building. b) At the third story, there was severe hinging at the top
of the columns (GPS coordinates: 0°57’20.1”S, 80°42’54.4”W).

8.7.3 Four-story residential building in Pedernales with commercial first floor. The 8-57
second story collapsed due to soft story (GPS coordinates: 0°4’7.3”N,
80°3’18.9”W).

8.7.4 Typical 2-story wood houses in Canoa. Some houses completely collapsed as 8-58
shown. (GPS coordinates: 0°27’46.9”S, 80°27’15.0”W)

8.7.5 Commercial reinforced concrete building. First, second and fourth floors 8-58
collapsed completely. (GPS coordinates: 1°3’20.3”S, 80°27’15.2”W).

8.7.6 Commercial reinforced concrete building in Portoviejo. Several pieces of the 8-59
facade collapsed (GPS coordinates: 1°3’18.9”S, 80°27’15.2”W).

8.7.7 Mall in Portoviejo with a spalled façade (GPS coordinates: 1°3’43.6”S, 8-59
80°27’55.3”W).

8.7.8 Partial collapse of nonstructural walls in governmental building in Portoviejo. 8-60


This building is occupied by SRI, the Service of Internal Rentals (GPS
coordinates: 1°2’60.0”S, 80°27’16.0”W).

8.7.9 Exterior façade and roof partial collapse in church. The church is located in the 8-60
central plaza in Pedernales along the Ruta del Spondylus (GPS coordinates:
0°4’12.4”N, 80°3’15.3”W).

8.8.1 Hospitals in Manabi and Santo Domingo de los Tsachilas. GEO Salud. 8-62
Ministerio de Salud Publica Ecuador. ©2013-2015 MSP
https://aplicaciones.msp.gob.ec/salud/publico/dniscg/geosalud/gui/#.

8.8.2 Hospitals in Manta City visited by the reconnaissance team. GEO Salud - 8-62
Ministerio de Salud Publica Ecuador ©2013-2015 MSP.
https://aplicaciones.msp.gob.ec/salud/publico/dniscg/geosalud/gui/#.

8.8.3 The 911 ECU center in Portoviejo. There was a large crack at the roof parapet 8-63
(GPS coordinates: 1°4’16.5”S, 80°26’47.5”W).

8.8.4 The Unity of Community Surveillance in Portoviejo. Minor spalling of 8-64


nonstructural exterior wall (GPS coordinates: 1°4’16.5”S, 80°26’47.5”W).

8.8.5 a) Unidad Educativa “Linus Pauling” in February 2015 (Google Maps ©2015 8-65
Google Inc.). b) Image taken during reconnaissance observation. The older
portion of this structure collapsed (GPS coordinates: 0°57’23.3”S, 80°42’41.0”W).

8.8.6 a) Side view of Unidad Educativa “Linus Pauling” in February 2015 (Google 8-66

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxx
Maps ©2015 Google Inc.). b) Reconnaissance photos of collapsed building after
the earthquake (GPS coordinates: 0°57’23.3”S, 80°42’41.0”W).

8.8.7 a) Nursery school and neighboring hotel building in Pedernales before the 8-67
earthquake. (Google Satellite Views, April 2015 Google Maps ©2015 Google Inc.).
b) The building next to the nursery school collapsed, was demolished and cleaned
out. (GPS coordinates: 0°4’11.7”N, 80°3’20.7”W).

8.8.8 a) Collapsed parapet along the exterior wall. b) Cracked nonstructural walls 8-68
(GPS coordinates: 0°57’14.9”S, 80°42’45”W).

8.8.9 a) Cracking and minor spalling in concrete nonstructural interior wall. b) 8-68
Cracking in masonry nonstructural exterior wall (GPS coordinates:
1°3’16.1S, 80°26’42.2”W).

8.8.10 Airport locations in the Manabí Area. Google Earth. .©2015. 8-69

8.8.11 Eloy Alfaro Airport, Manta. Google Earth.©2015. 8-70

8.8.12 Evacuated International Airport Eloy Alfaro (GPS coordinates: 0°57’10.7”S, 8-71
80°41’3.2”W).

8.8.13 Beam to column connection failure. The rebar at the column is slightly bent 8-71
and there is significant spalling present (GPS coordinates: 0°57’11.7”S,
80°41’3.6”W).

8.8.14 Cracked and collapsed nonstructural brick walls in the International Airport 8-72
Eloy Alfaro (GPS coordinates: 0°57’10.7”S, 80°41’3.2”W).

8.8.15 Partial collapse of a nonstructural wall. There were dowels from the columns 8-72
embedded in the wall. (GPS coordinates: 0°57’10.7”S, 80°41’3.2”W).

8.8.16 a) International Airport Eloy Alfaro Control Tower (Google Satellite View, 8-73
February 2015 Google Maps ©2015 Google Inc.) b) The location of the control
tower which has been cleaned out after collapse (GPS coordinates:
0°57’9.5”S, 80°41’1.9”W).

8.8.17 Los Perales Airport, San Vicente. Google Earth ©2015. 8-74

8.8.18 Pounding and collapse of parapet at the Airport of the Perales (GPS 8-75
coordinates: 0°36’5.3”S, 80°24’27.0”W).

8.8.19 Pounding of columns and roof parapet at the Airport of the Perales terminal 8-75
(GPS coordinates: 0°36’5.3”S, 80°24’27.0”W).

8.8.20 The newer portion of the runway at the Airport of the Perales had cracks along 8-76
the runway (GPS coordinates: 0°36’58.7”S, 80°23’53.8”W).

8.10.1.1 Bridge site looking eastward from San Vicente to Bahía de Caráquez 8-76
(GPS coordinates: 0°36'30.66"S ,80°24'30.59"W).

8.10.1.2 Plan of view of Los Caras Bridge, Courtesy of Ecuador’s Army Corps of 8-77
Engineering (GPS coordinates: 0°36'33.58"S, 80°24'58.68"W).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxxi
8.10.1.3 Bridge cross-section (Courtesy of Ecuador’s Army Corps of Engineers). 8-78

8.10.1.4 Soil profile Los Caras Bridge (Courtesy of Ecuador’s Army Corps of 8-79
Engineers and Adolfo Caicedo).

8.10.1.5 Ecuador’s Army Corps of Engineers team shows the document inspection of 8-80
MIDUVI (GPS coordinates: 0°36'48.85"S, 80°24'46.93"W).

8.10.1.6 Modular bridge long expansion joint (GPS coordinates: 0°36'48.85"S, 8-80
80°24'46.93"W).

8.10.1.7 Modular bridge long expansion joint, San Vincente side (GPS coordinates: 8-81
0°36'30.74"S, 80°24'30.54"W).

8.10.1.8 Neoprene Rubber Standard Expansion joint (GPS coordinates: 8-81


0°36'35.28"S, 80°25'17.15"W).

8.10.1.9 External movements synchronizers (GPS coordinates: 0°36'33.58"S, 8-81


80°24'58.68"W).

8.10.1.10 Cracking of the pier 3 on the horizontal shear block at the Bahia side, the 8-82
pier has elastomeric bearings (GPS coordinates: 0°36'36.56"S and 80°25'27.26"W).

8.10.1.11 Friction pendulum device as part of the base isolation system of Los Caras 8-82
Bridge (Left), Inner displacement isolator view (GPS coordinates: 0°36'33.58"S,
80°24'58.68"W).

8.10.3.1 IESS Hospital Location, aerial view, Google Earth 2016. (GPS coordinates: 8-84
0°57'17.90"S, 80°43'26.70"W).

8.10.3.2 Front façade of the IESS Hospital in Manta (GPS coordinates: 8-84
0°57'17.90"S, 80°43'26.70"W).

8.10.3.3 Rebar exposure and cold joint presence in beams of the IESS Hospital in 8-85
Manta (GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

8.10.3.4 Damage above columns on the 2-story portion of the IESS Hospital in 8-86
Manta (GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

8.10.3.5 Cracks in or near the beam-column joints at the 2nd story of the 5-story 8-86
portion of the IESS Hospital in Manta (GPS coordinates: 0°57'17.90"S,
80°43'26.70"W).

8.10.3.6 Staircase state from the 2nd to 5th story of the 5-story portion of the 8-87
IESS Hospital in Manta (GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

8.10.3.7 Appearance of concrete bricks used in the façade of the IESS Hospital in 8-88
Manta (GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

8.10.3.8 Cross tension cracks and spalling in hollow clay interior wall in the IESS 8-88
Hospital in Manta (GPS coordinates: 0°57'17.3"S, 80°43'25.2"W).

8.10.3.9 a) Collapse of non-structural concrete block exterior walls in the IESS 8-88
Hospital in Manta. The interior nonstructural walls are clearly not full-height and

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxxii
are not braced horizontally (GPS coordinates: 0°57'18.1"S, 80°43'25.3"W).

8.10.3.10 Fallen parts of the ceiling in the IESS Hospital in Manta (GPS 8-89
coordinates: 0°57'18.1"S, 80°43'25.3"W).

8.10.3.11 Fallen ducts and ceilings in the IESS Hospital in Manta (GPS 8-89
coordinates: 0°57'18.1"S, 80°43'25.3"W).

8.10.3.12 Front façade of the Rodríguez Zambrano Hospital in Manta (GPS 8-90
coordinates: 0°57'14.2"S, 80°4330.9"W).

8.10.3.13 Reinforced concrete construction with moment resisting concrete frames 8-91
with waffle slabs in the Rodríguez Zambrano Hospital in Manta (GPS
coordinates: 0°57'21.8"S, 80°44'30.8"W).

8.10.3.14 Cross tension cracks in interior nonstructural walls in the Rodríguez 8-92
Zambrano Hospital in Manta (GPS coordinates: 0°57'15.8"S, 80°44'32.6"W).

8.10.3.15 Spalling in hollow clay interior walls in the Rodríguez Zambrano Hospital in 8-92
Manta (GPS coordinates: 0°57'19.3"S, 80°44'31.53"W).

8.10.3.16 The highlighted area used to have nonstructural walls that collapsed during 8-93
the earthquake. This is the 6th floor in the Rodríguez Zambrano Hospital in Manta.
Refer to Section 11.6 (GPS coordinates: 0°57'14.9”S, 80°44'31.9"W).

8.10.3.17 Medical equipment in the Rodríguez Zambrano Hospital in Manta. a) A 8-94


large lighting equipment was well anchored into the ceiling and did not fall during
the earthquake. b) Tanks in this hospital room were chained to the wall so that they
would not fall (GPS coordinates: 0°57'21.8"S, 80°44'30.7"W).

8.10.3.18 Structural and non-structural damage Miguel H. Alcívar hospital during the 8-95
1998 earthquake in Ecuador (Courtesy of Aguiar R.).

8.10.3.19 Miguel H. Alcívar Hospital (GPS coordinates: 0°37'18.37"S, 80°25'39.52"W). 8-95

8.10.3.20 Non-structural damage: Masonry infill - Outpatient (GPS coordinates: 8-96


0°37'18.46"S, 80°25'38.62"W).

8.10.3.21 Non-structural damage: Masonry infill- Outpatient (GPS coordinates: 8-96


0°37'18.20"S, 80°25'38.95"W).

8.10.3.22 Non-structural damage contents: Ground level (GPS coordinates: 8-97


0°37'18.05"S, 80°25'39.00"W).

9.1.1 Map depicting the main cities of Ecuador (photo from web, gosouthamerica.com). 9-1

9.2.1 Emergency personnel and volunteers trying to save a person trapped in car in Guayaquil 9-2
right after the earthquake on April 16th. Photo: R. Cedeño (Vistazo Magazine, 2016).

9.2.2 Volunteers search for survivors in the debris of buildings in Pedernales on April 17th. 9-3
Photo by US Dept. of State (blogs.state.gov/stories/2016/04/17/earthquake-ecuador).

9.2.3 Army personnel deployed with helicopters to affected areas on April 17th. Photo: 9-3
Ecuadorian Ministry of Defense (Dialogo Americas, 2016).

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxxiii
9.2.4 Zones assigned to Rapid Response Teams (SENPLADES, 2012 9-4

9.2.5 CRE volunteer providing pre-hospital care and implementing family links restoration 9-5
program available for people affected by the earthquake. Photo by CRE.

9.2.6 A woman and her daughter are rescued Sunday after they were trapped in the rubble 9-6
from the earthquake. Photo by National Police of Ecuador (NBC, 2016).

9.2.7 Moment of silence observed by rescue and recovery team in Pedernales (4/19/16, ~GPS 9-7
0o5'0''N, 80o7'0''W). Photo provided by the Ecuadorian Armed Forces.

9.2.8 Army rescue and recovery team in Pedernales (4/19/16, ~GPS 0 o5'0''N, 80o7'0''W). 9-7
Photo provided to GEER-ATC team by the Ecuadorian Armed Forces.

9.3.1 Army cleanup operations in Carapoto, Manabí (4/21/16, ~GPS 0 o5'0''S, 80o29'0''W). 9-8
Photo provided to GEER-ATC team by the Ecuadorian Armed Forces.

9.4.1 UNHCR (United Nations Refugee Agency) staff along with police and local people 9-9
work install tents in Chamanga, Esmeraldas (approx. GPS 0 o16'0''N, 79o56'0''W).
Photo by UNHCR/Santiago Arcos.

9.5.1 Treating victims outdoors in R. Rodriguez Zambrano Hospital (Teleamazonas, 2016). 9-11

9.5.2 Volunteers and MIES members distribute food supplies (top). Members of the National 9-12
Police organize water bottles in Pedernales on April 20th (bottom). Photo by M. Ayala,
(Andes, 2016).

9.5.3 Potable water distribution on April 20, 2016. Photo by the Ecuadorian Armed Forces. 9-14

9.5.4 Belen Carillo, a UNICEF child protection specialist, with Kimberly, 6, at the “Y” 9-14
shelter that hosted 250 displaced families outside of San Jose de Chamanga. Web
photo UNICEFUSA/Sandler.

9.5.5 A young boy, potentially at risk for post-disaster trauma, colors in a child-friendly 9-15
space in Portoviejo. Photo 4/19 UNICEF/ECU/Castellanos (unicefusa.org/ecuador).

9.5.6 Pedernales, April 24: Kiara Farias, 2, whose arm was broken in the earthquake plays 9-16
with her brother Jostin, 6, in a makeshift camp for displaced people (top). Volunteers
serve food to children (bottom). Photos by R. Abd, AP (fresnobee.com/article74243762).

9.6.1 An air force soldier stands guards the area in front of collapsed buildings in Manta. 9-17
Photo by R. Abd, AP (4/17/16, approximate GPS: 0°57′43″S, 80°42′45″W).

9.6.2 Demolition and debris removal in Portovejo. Photo by GEER-ATC member R. Gilsanz 9-18
(4/28/16, GPS: 1°3'24.61"S, 80°27'17.29"W).

9.6.3 Army cleanup and restoration operations for road infrastructure. Photo provided to 9-18
GEER-ATC team by the Ecuadorian Armed Forces.

9.7.1 Damaged propane tanks in Pedernales home. Photo by the Ecuadorian Armed Forces. 9-19

9.7.2 Fuel Relief Fund propane distribution with local vehicles and volunteers (FRF, 2016). 9-20

9.7.3 Volunteers distribute propane tanks for cleaning the water and cooking (FRF, 2016). 9-20

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxxiv
9.7.4 Shelter in Pedernales after gas explosion (El Comercio, 2016). 9-21

9.7.5 House affected by a fire in Portoviejo 2 days after the earthquake (El Universo, 2016). 9-22

9.7.6 Burned belongings in a Portoviejo street 2 days after the event (El Universo, 2016). 9-22

9.7.7 The fire seen in one of the streets of Tosagua (El Telégrafo, 2016; El Diario, 2016). 9-22

GEER-ATC Muisne, Ecuador, Earthquake List of Figures


Report Version 1 xxxv
LIST OF TABLES
Table Page

2.1.1 Ecuador provinces and their capitals (Wikipedia, 2016). 2-1


2.1.2 Ecuador coastal lowlands provinces: population and area characteristics (ACAPS, 2016). 2-5

2.1.3 Ecuador coastal lowlands main parroquias (parishes): population, areas (ACAPS, 2016). 2-5
2.1.4 Ecuador demographic profile (ACAPS, 2016). 2-6
2.2.1 GEER/ATC Team Reconnaissance Focus. 2-10
lowlands region of Ecuador (ACAPS, 2016).

4.1.1 Active volcanoes of continental Ecuador and max VEI (Smithsonian Institute’s Global 4-8
Volcanism Program, volcano.si.edu/world/region.cfm?rnum=1502, and Toulkeridis,
(2013).

4.2.1 Historic events with M > 6 from 1556 to 2016 (NOOA earthquake database, 2016). 4-16

4.2.2 Description NOAA database index for death, injuries, economical and house losses. 4-18

5.1.1 Seismological parameters for the main 4/16/16 event by IG-EPN, USGS, and CMT. 5-1

5.1.2 Seismological parameters for major aftershocks of the 4/16/16 event (IG-EPN, 2016). 5-3

5.2.1 Accelerometer stations in Ecuador owned by RENAC (maintained by IG-EPN, 5-6


Singaucho et al., 2016), OCP, and LMI with PGA values recorded during the main event.

5.2.2 Summary of 10 of the stations with ground motion records processed by GEER. 5-8

5.2.3 Parameters for the BCHydro GMPE (Abrahamson et al., 2016). Stations were assumed 5-40
to be in the fore-arc, although some far stations are in the back-arc, east of Andeans
mountains.

5.2.4 Parameters for constructing the NEC-15 design acceleration response spectra for 5-46
stations APED, PDNS, AMNT, APO1, ACHN, and AGY2.

7.2.1 Geotechnical information pre- and post-earthquake during reconnaissance. 7-6

7.2.2 Manta Port: Pre-earthquake boring coordinates. 7-9

7.2.3 Mejia Bridge: Pre-earthquake boring coordinates. 7-10

7.2.4 Las Chacras Bridge: Pre-earthquake boring coordinates. 7-13

7.2.5 Briceño Bridge: Pre-earthquake boring coordinates. 7-14

7.2.6 Chone Dam: Pre-earthquake boring coordinates. . 7-16

7.2.7 Manta Port parking area: In-situ measured Vs profile. 7-19

7.2.8 Manta Port wharf deck: Median in-situ measured Vs profile. 7-22

7.2.9 Las Chacras (Rio Chico) Bridge: Median in-situ measured Vs profile. 7-25

GEER-ATC Muisne, Ecuador, Earthquake List of Tables


Report Version 1 xxxvi
7.2.10 Chone Dam: Median in-situ measured Vs profile. 7-27

7.2.11 Middle of Chone Dam: Median in-situ measured Vs profile. 7-28

7.2.12 Manta SMS: Median in-situ measured Vs profile. 7-36

7.2.13 Central Manta: Summary of measured HVSR peaks. 7-37

7.2.14 Median Vs profile at IESS Hospital. 7-40

7.3.1 Manta AMNT station: Median in-situ measured Vs profile. 7-45

7.3.2 Characteristics of five local seismic events recorded at AGYE2, near Boring ERU. 7-58

7.5.1 Retaining structures details and damage. Actual wall heights could not be confirmed as 7-79
they were partially buried; for Mejia Bridge, heights are based on design drawings.

7.9.1 Soil description from two borings at Rio Chico (Las Chacras) Bridge embankments. 7-165

7.10.1 Median Vs values measured at IESS Hospital. 7-177

8.3.1 Factors Z as a function of the seismic zones in the country. 8-17

8.3.2 Factors S as a function of site conditions. 8-18

9.2.1 Emergency response personnel deployed one day after the earthquake (SGR-14, 9-5
2016). Referenced agencies: (*) EDAN: Damage Assessment & Needs Analysis
(Evaluación de Daños y Análisis de Necesidades) and (**) SSC: Rural Social Security
(Seguro Social Campesino).

9.2.2 Emergency response volunteers deployed 1 day after the earthquake (SGR-14, 2016). 9-6

9.4.1 Shelters activated immediately after the earthquake (SGR Status Report 14, 2016). 9-9

9.4.2 MIES emergency response teams (left) and tents distributed (right) after SGR Status 9-10
Status Report 14 (2016).

9.4.3 Active shelters a week after the earthquake (SGR Status Report 41, 2016). 9-10

9.4.4 Active shelters a month after the earthquake (SGR Status Report 41, 2016). 9-11

9.5.1 Kits sent to Manabi in response to the Emergency (SGR Status Report 14, 2016). 9-13

9.5.2 Kits sent to Esmeraldas in response to the Emergency (SGR Report 14, 2016). 9-13

9.5.3 Kits sent to affected areas in response to the Emergency (SGR Report 14, 2016). 9-13

9.5.4 Food kits from 4/17 to 5/17 provided by MIES and the Armed Forces. 9-15

GEER-ATC Muisne, Ecuador, Earthquake List of Tables


Report Version 1 xxxvii
Blank Page
Blank Page
CHAPTER 1
Introduction

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
1 Introduction
On the evening of April 16th, 2016, a moment magnitude Mw7.8 earthquake struck northern
Ecuador, offshore of its west coast. The earthquake was named Muisne after the city of its
epicenter, located about 29 km south-southeast of the town of Muisne, in the province of
Manabí, at a hypocentral depth of 21 km. In the first 24 hours, over 135 aftershocks were
recorded with hundreds more in the weeks that followed. Two major aftershocks with Mw6.7
and 6.9 occurred in the early morning of May 18, 2016, about 30 km from the epicenter of the
main event and destroyed areas in the north that were already heavily damaged.

Ecuador has a remarkable seismic history and a unique mixed tectonic setting of a
subduction zone with crustal earthquakes and volcanism, and difficult subsurface conditions
that can either generate large amplification of the earthquake waves propagating from the
bedrock, or liquefy under strong shaking. The April 16th event was caused by shallow thrust
faulting on or near the boundary where the Nazca plate subducts beneath the S. American plate
61 mm/yr according to the United States Geological Survey (USGS). Subduction along the
Ecuador and the Peru-Chile trenches has led to uplift of the Andes mountains and has produced
destructive earthquakes, including the largest event ever recorded, the 1960 Mw9.5 southern
Chile earthquake. Historic earthquake activity of events with estimated magnitude greater than
7.5 traces back to 1556. The mechanism of the 2016 Muisne earthquake mostly relates to events
of the last century that started with the 1906 megathrust subduction Esmeraldas earthquake
with estimated magnitude of 8.8 and a rupture area that captured sub-areas of earthquakes that
followed, including the 2016 one that closed a seismic gap observed since 1906.

Overall, the April 16th Muisne earthquake and its aftershocks led to hundreds of fatalities,
thousands of injuries, tens of thousands homeless and an economic impact estimated at 3.5%
of the nation’s Gross Domestic Product (GDP). Due to its severe impact on the nation’s
infrastructure, this event has been classified as almost equally damaging as the 2010 Haiti
earthquake. The event imposed a detrimental shock to the health infrastructure with
incapacitation of 10 hospitals and 15 health centers servicing 1,500 beds, due to structural or
nonstructural damage with direct accumulated losses estimated at 150M$US. It demonstrated
extreme ground motions with prominent site amplification and liquefaction effects, but also
resilient behavior of the seismically isolated Los Caras Bridge, which sustained the largest ever
recorded bearing displacement beyond design levels, and kept the bridge safe and functional.

GEER-ATC Muisne, Ecuador, Earthquake 1-Introduction


Report Version 1 1-1
The reconnaissance mission for the 2016 Muisne earthquake was unique for two main
reasons. First, it brought together the local earthquake engineering community and the GEER-
ATC group, to form a multidisciplinary team of over 40 people, who documented geotechnical,
structural and nonstructural effects. Observations included earthquake-induced liquefaction,
soil amplification and settlement effects, and failures of essential buildings as well as
successful behavior of structures with seismic protection design.

Second, the members of this mission had the opportunity to discuss with Ecuadorian
government officials immediate and long-term rebuild planning after this national disaster.
Information was brought back to USA and sparked the interest of our government officials to
discuss the opportunities this earthquake has presented to incorporate multi-hazard resiliency
concepts in rebuilding. A unique, direct dialogue between geotechnical and structural engineers
with the highest levels of our government continued on the Ecuador needs and the ability of
US to help in: (i) understanding the natural hazards exposure; (ii) transferring knowledge and
using seismic protective technologies; and (iii) set resiliency goals to address short-term needs
of accommodating homeless families and offering health care, but also long-term, risk-based
planning to be able to anticipate, respond and recover from future earthquakes and other
hazards such as tsunamis, landslides, infectious diseases, and climate change.

The reconnaissance has yielded datasets, lessons, and suggestions for future research on:
(i) seismological and recorded strong ground motions
(ii) application of US rapid assessment methods
(iii) geotechnical and infrastructure performance
(iv) site amplification and liquefaction on the recorded motions
(v) structural and nonstructural effects and correlation to geotechnical effects
(vi) critical facilities and seismically-isolated structures response
(vii) use of advanced 3-D technologies to map drone and photo observations
(viii) case histories documentation of successful behavior in addition to failures
(ix) community/government response and awareness aspects.

In future research, the collected data, including ground motion records and geotechnical
and structural documentation can be used, in combination with much-needed in-situ
geotechnical testing and 3-D mapping to further understand the observations and analyze the
good and bad behavior of the natural and built environment. Other needs include evaluation
and testing of model houses that can be replicated fast in the field to house the homeless, and
development of simple guidelines and procedures that can be used to improve the seismic
performance of health care, transportation, and critical facilities in the future.
GEER-ATC Muisne, Ecuador, Earthquake 1-Introduction
Report Version 1 1-2
Blank Page
Blank Page
CHAPTER 2
Ecuador and Reconnaissance

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
2.1 Ecuador
Ecuador is a country in north-western South America, bordered by Colombia to the north,
Peru to the east and south, and Pacific Ocean to the west (Fig. 2.1.1). It was named after its
geographic location straddling the Equator. Ecuador has an area of 283,560 km2, including the
famous Galápagos Islands, and is comprised of various ethnic groups and ecosystems living
and natural landscapes. There are 4 regions divided in 24 provinces, including the Coastal
Lowland, Sierra (Andean Highlands), Amazon Rainforest (Oriente), and the Galápagos. Each
region has a distinct climate, geology, and culture. The provinces and regions comprising
Ecuador are shown in Fig. 2.1.2. Each province is divided into cantons, which are in turn
divided into parroquias or parishes. Table 2.1.1 provides information on the provinces.

Figure 2.1.1. Ecuador Geography: Location in South America. Base image: World Atlas (©2016).

The capital city of Ecuador is Quito (Fig. 2.1.3) located at elevation 2,850 m above sea
level, and its largest city is Guayaquil (Fig. 2.1.4). Parts of Quito, especially the old city, have
been designated a cultural site of “Outstanding Universal Value” by the United Nations
Educational Scientific & Cultural Organization (UNESCO). World Heritage Center. UNESCO
(2016) states that “The city of Quito has preserved many of its attributes that make it a city of
Outstanding Universal Value. The Historic Centre of Quito has conserved its original
architecture and form, with new constructions being built outside of the colonial center. The
original design of the city by Dionisio Alcedo y Herrera, which dates back to 1734, is still
clearly visible. The city conserves the least modified historic center of all Latin America.”
GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance
Report Version 1 2-1
Figure 2.1.2. Ecuador provinces (top, Wikipedia, 2016) and regions (Your Escape to Ecuador, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-2
Table 2.1.1. Ecuador provinces and their capitals (Wikipedia, 2016).
Map Population as per the Area
Province Capital
key census on 11/28/2011 (km²)
1 Azuay Cuenca 712,127 8,309.58
2 Bolívar Guaranda 183,641 3,945.38
Map Population as per the Area
3 Province
Cañar Capital
Azogues 225,184 3,146.08
key census on 11/28/2011 (km²)
4
1 Carchi
Azuay Tulcán
Cuenca 164,524
712,127 3,780.45
8,309.58
2
5 Bolívar
Chimborazo Guaranda
Riobamba 183,641
458,581 3,945.38
6,499.72
3
6 Cañar
Cotopaxi Azogues
Latacunga 225,184
409,205 3,146.08
6,108.23
4 Carchi Tulcán 164,524 3,780.45
7 El Oro Machala 600,659 5,766.68
5 Chimborazo Riobamba 458,581 6,499.72
8 Esmeraldas Esmeraldas 534,092 16,132.23
6 Cotopaxi Latacunga 409,205 6,108.23
7 El Oro Puerto Baquerizo
Machala 600,659 5,766.68
9 Galápagos 25,124 8,010.00
Moreno
8 Esmeraldas Esmeraldas 534,092 16,132.23
10 Guayas Guayaquil 3,645,483 15,430.40
Puerto Baquerizo
9 Galápagos 25,124 8,010.00
11 Imbabura Moreno
Ibarra 398,244 4,587.51
10
12 Guayas
Loja Guayaquil
Loja 3,645,483
448,966 15,430.40
11,062.73
11 Imbabura Ibarra 398,244 4,587.51
13 Los Ríos Babahoyo 778,115 7,205.27
12 Loja Loja 448,966 11,062.73
14 Manabí Portoviejo 1,369,780 18,939.60
13 Los Ríos Babahoyo 778,115 7,205.27
15
14
Morona Santiago
Manabí
Macas
Portoviejo
147,940
1,369,780
24,059.40
18,939.60
16
15 Napo
Morona Santiago Tena
Macas 103,697
147,940 12,542.50
24,059.40
16 Napo Puerto Francisco
Tena de 103,697 12,542.50
17 Orellana 136,396 21,692.10
Orellana
Puerto Francisco de
17 Orellana 136,396 21,692.10
18 Pastaza Orellana
Puyo 83,933 29,641.37
18 Pastaza Puyo 83,933 29,641.37
19 Pichincha Quito 2,576,287 9,535.91
19 Pichincha Quito 2,576,287 9,535.91
20 Santa Elena Santa Elena 308,693 3,690.17
20 Santa Elena Santa Elena 308,693 3,690.17
Santo Domingo de Santo Domingo de
21
21
Santo Domingo de Santo Domingo de 368,013
368,013
3,446.65
3,446.65
los
los Tsáchilas
Tsáchilas los
los Colorados
Colorados
22
22 Sucumbíos
Sucumbíos Nueva Loja
Nueva Loja 176,472
176,472 18,084.42
18,084.42
23 Tungurahua Ambato 504,583 3,386.25
10,584.26
10,584.26
24
24 Zamora-Chinchipe
Zamora-Chinchipe Zamora
Zamora 91,376
91,376

Ecuador is known for its diverse ecosystem and unique fauna and flora. It has 2,237 km of
coastline and an array of climate ranging from tropical along the coastline, to a cooler more
arid atmosphere in the inlands and high mountains, and a rainforest climate in the Amazon
basin. The existence of various climates is the reason that Ecuador’s economy consists mostly
of agriculture and fishing, with tourism increasing recently; nevertheless, the country’s main
export is crude oil. Shrimp farming is another major source of income for many local families,
with shrimp farming covering more than 95% of Ecuadorian aquaculture according to the
United Nations Food and Agricultural Organization (FAO, 2005).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-3
Figure 2.1.3. Quito. Base image from Lifewithoutlimbs (©2013).

Figure 2.1.4. City of Guayaquil. Base image: Municipal GAD Guayaquil (©2014).

Ecuador was inhabited by a various Amerindian groups that became part of the Inca Empire
during the 15th century. The country was colonized by Spain during the 16 th century when
Quito became a seat of Spanish colonial government in 1563 and part of the Viceroyalty of
New Granada in 1717 (CIA, 2016). It gained independence in 1820 by joining the Bolivarian

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-4
Union of Colombia, Venezuela and Panama, and became fully independent in 1830. The rich
and diverse Ecuadorian history is reflected in its population that is over 52% of Native
American ethnicity, 42% of European ancestry, and 6% of African origins (Nature
Communications, 2015). Overall, most of the population has a multi-ethnic background.

Ecuador has the highest population density in South America. Similar to other Andean
countries, it has experienced a population increase, the result of both a decreasing death rate
and a continued high birth rate (McLeod et al., 2016). The population is concentrated in major
urban areas and along the coastline where fishing and farming thrive. Additionally, Ecuador is
host to many immigrants, mainly from neighboring Colombia and Peru; seeking refuge from
armed conflict and drug-related violence, or pursuing better economic opportunities. Tables
2.1.2 and 2.1.3 present the population distribution along Ecuador’s coastal region, as well as
the population density per km2. These regions are densely populated.

Table 2.1.2. Ecuador coastal lowlands provinces: population and area characteristics (ACAPS, 2016).
Santo
Manabi Esmeraldas Los Rios Guayas Santa Elena
Domingo

Total population 1,369,780 534,092 778,115 368,013 3,645,483 308,693

Population density
72 33 108 107 236 84
(per km2)
Santo
State capital Portoviejo Esmeraldas Valdivia Guayaquil Santa Elena
Domingo

Table 2.1.3. Ecuador coastal lowlands main parroquias (parishes): population, areas (ACAPS, 2016).

Pedernales Portoviejo Manta Muisne

Total population 33,640 223,086 221,122 8,880

Population in urban
65% 93% 98% 67%
areas
Population density
2 44 534 1,046 50
(per km )

About 64% of the country’s population live in urban areas (CIA, 2016) with demographic
profile shown on Table 2.1.4. Among the 0-14 age group, 25% of the children are chronically
malnourished (International Food Policy Research Institute, 2015) with mortality rate
relatively higher than that of neighboring countries. The infant mortality rate is 18 every 1,000
live births, increasing to 22 for the under-5 group (World Bank, 2015).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-5
Table 2.1.4. Ecuador demographic profile (ACAPS, 2016).

Age (years) Portion of Population Male Female

0-14 28% 51% 49%

15-24 19% 51% 49%

25-54 39% 49% 51%

55-64 7% 49% 51%

65 + 7% 48% 52%

Large populations, especially those in urban areas, are very conducive to the spread of
infectious diseases. About 0.34% of adults are infected with HIV. The most common high-risk
diseases are bacterial diarrhea, hepatitis A, typhoid fever, dengue fever and malaria (CIA
Factbook, 2013). Due to its geographical location, Ecuador is prone to the Zika epidemic,
although there are only 140 cases of infected people by the virus as of April 2016 and some
recent reports in Portoviejo and Guayaquil (El Comercio, 2016).

Historically, Ecuador has contributed significantly to scientific research, and the literacy
rate in the country is relatively high with 94.5% of citizens ages 15 and over able to read and
write. According to the CIA (2015) data, 95.4% of males and 93.5% of females are literate in
Ecuador, with a school life expectancy of 14 years for males and 15 years for females. The first
scientific expedition to measure the circumference of the Earth, led by Charles-Marie de La
Condamine of France, was based in Ecuador. Charles Darwin did abundant research to
establish basic theories of modern geography, ecology, and evolutionary biology as
documented his renowned publication “The Origin of Species” (McLeod et al., 2016).

Figure 2.1.5. Sea turtles off the coast of the Galápagos Islands (Avalon Waterways, ©2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-6
Located in the Pacific Ocean, some 1,000 km west of the mainland, the remoteness and
uniqueness of the Galápagos proved a fertile ground for evolution studies due to a vast number
of endemic species that inhabit the islands (Fig. 2.1.5). With an archipelago of 19 islands and
population of 25,000 (National Institute for Statistics & Census of Ecuador, 2010), the
Galápagos are as pristine and unspoiled by man as they are beautiful and remarkable (Fig.
2.1.6). This setting allowed for unique evolutionary developments in species not found
anywhere else, making Galápagos a designated World Heritage site in 1978 (UNESCO, 2016).

Figure 2.1.6. Sunset at Galápagos Islands in the Pacific Ocean (Green Coconut Rum, ©2016).

Owing to its colonial past, Spanish is the official language of Ecuador spoken by the
majority of the population, as is the case with almost all South American countries. Still many
local languages and dialects persist, especially in rural areas and villages like Quechua and
Shuar that experience income inequality and poverty, mostly affecting indigenous groups. Also
typical of the region, Ecuador is an overwhelmingly Roman Catholic country with the church
having a significant role in its society. According to national census data, 80.4% of the
population is Roman Catholic, 11.3% Evangelical Protestants, 1.29% Jehovah's Witnesses, and
7% other, mainly Jewish, Buddhists and Latter-day Saints (El Universo, 2012).

Ecuador’s government is a presidential republic of three branches: executive, legislative,


and judicial with the president being head of state and government. The current president is
Rafael Correa Delgado, since January of 2007 when he assumed this office. The president
appoints members of his cabinet and is commander-in-chief of the Ecuadorian armed forces,
comprised of the Land Force (Fuerza Terrestre Ecuatoriana, FTE), Ecuadorian Navy (Fuerza

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-7
Naval del Ecuador (FNE) which includes the Naval Infantry and Aviation, Coast Guard, and
Air Force (Fuerza Aerea Ecuatoriana, FAE). The legislative branch consists of a unicameral
National Assembly (Fig. 2.1.7). The judicial branch is formed by the Superior Courts and
Tribunals and the National Council of the Judiciary (Organization of American States, 2011).

Figure 2.1.7. Republic of Ecuador National Assembly. Base image: La Info (©2014).

The Ecuadorian economy uses US$ as its currency and has been growing over the past
few years. From 2006 to 2014, the Gross Domestic Product (GDP) growth averaged 4.6%,
thanks to robust oil prices and external financing flows (World Bank, 2016). In 2015,
Ecuador’s GDP was $100.9 billion. This stimulus allowed for increased social spending,
particularly in the sectors of energy and transportation. Ecuador relies heavily increases of oil
prices for increased government revenue.

The extreme poverty rate has declined significantly between 1999 and 2010 (Centre for
Economic and Policy Research, 2012). In 2001 it was estimated at 40% of the population,
dropping to 17.4% by 2011 (Ray & Kozameh, 2012). Nevertheless, the situation reversed as
the price of a barrel of oil dropped from $100 to $30 in the last couple of years. This, coupled
with the international financial crisis, severely affected the economy of the country. The Mw7.8
April 16th, 2016 earthquake generated an additional and significant negative impact in the
economy of Ecuador.

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-8
2.2 Reconnaissance Focus
The epicenter of the Mw7.8 April 16, 2016 earthquake was in the northern coast of Ecuador,
about 30 km SE of Muisne in Esmeraldas. The damage was most intense south of the epicenter
and down the western coast, mainly in the province of Manabí. Reconnaissance was focused
in this area, extending roughly 140 km from Manta to Pedernales (the closest town to the
epicenter). The focal points of reconnaissance on the ground are mapped on Fig. 2.2.1 and
listed on Table 2.2.1. In addition, we spent several hours in a military helicopter that flew us
over the affected areas. During the flyover we observed the 4 dams of Table 2.2.1 that were
inspected by local engineers. Our team observed one of these facilities, the Chone dam.

Figure 2.2.1. GEER-ATC team reconnaissance on the ground with focus cities shown with red pins.
Dams observed by military helicopter flyover shown as white pins. Base Image: Google Earth ( ©2016).

The towns visited by the GEER-ATC team include Portoviejo, which is the capital of the
Manabí province, and as such, a political and economic center. It is also the most populated
city in Manabí with about 280,000 people, followed by Manta with population of about
220,000. The rest of the cities visited are on the coast with average populations of about 20,000,
and some of which being popular tourist destinations like Bahia de Caraquez.

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-9
Table 2.2.1. GEER-ATC Team Reconnaissance Focus.
Reconnaissance Focus Geographic Coordinates (degrees)
Pedernales 0° 4' 21.1" N 80° 03' 5.0" W
Jama 0° 12' 14.8" N 80° 15' 49.1" W
Canoa 0° 28' 1.6" N 80° 27' 12.7" W
Bahia de Caraquez 0° 37' 19.5" N 80° 25' 38.9" W
Rocafuerte 0° 55' 29.5" N 80° 27' 7.5" W
Jaramijó 0° 57' 25.3" N 80° 38' 18.5" W
Manta 0° 58' 7.0" N 80° 42' 32.0" W
Portoviejo 1° 3' 19.4" S 80° 27' 9.1" W
Chone/Rio Grande Dam 0° 41' 55.5" N 79° 59' 19.2" W
La Esperanza Dam 0° 53' 18.8" N 80° 4' 34.1" W
Poza Honda Dam 1° 6' 49.4" S 80° 12' 13.5" W
San Vicente Dam 2° 0' 21.6" S 80° 32' 1.4" W

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-10
2.3 Reconnaissance Routes & Objectives
During the main reconnaissance period in Ecuador, from April 26 th to May 1 st, the GEER-
ATC teams was divided in sub-teams with a leader of each sub-team. The division in sub-teams
was based on geographic and technical criteria. Since the team included both geotechnical and
structural engineers, we often followed up a structural site of interest reconnaissance with a
geotechnical team and vice versa. Each sub-team included at least one engineer of each
discipline (e.g., a geotechnical sub-team included at least one structural team member), to
enhance our observations and have contribution from different specialties. This section
presents the sub-teams with their GPS routes and objectives for each day of reconnaissance.

DAY 1 - APRIL 26, 2016 (MANTA)


The first day, the team focused on collecting existing information for planning the
reconnaissance targets with the help of the Ministry of Housing. Our Ecuadorian partners
addressed logistics of transportation and priorities and provided IDs for each US team member
from the governmental security agency ECU911 that allowed for access to damaged sites.

DAY 2 - APRIL 27, 2016 (MANTA)


Team 1: C. Wood, I. Ochoa of Geoestudios, G. Lyvers.
Objectives: Shear wave velocity Vs measurements and Horizontal-to-Vertical Spectral Ratio
(HVSR) testing at the Manta AMNT strong motion station and port.
Route: GPS track and locations are shown on Fig. 2.3.1.

Figure 2.3.1. Team 1, 4/27/16: GPS track for measurements at Manta strong motion station and port.
Base Image: Google Earth (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-11
Teams 2, 3: S. Nikolaou, A. Zekkos, K. Rollins, D. Alzamora, A. Caicedo, G. Ponce.
Objective: Manifestation of liquefaction observations and measurements at Port of Manta.
Route: GPS track is shown on Fig. 2.3.2 (including helicopter flyover).

Figure 2.3.2. Teams 2, 3, 4/27/16: GPS track at Manta Port. Base Image: Google Earth (©2016).

Team 4: R. Gilsanz, V. Diaz, R. Luque, C. Arteta.


Objective: Building damage survey in Manta and correlation to liquefaction observations.
Route: GPS track is shown on Fig. 2.3.3.

Figure 2.3.3. Team 4, 4/27/16: GPS track of structural damage survey at Manta. Base Image: Google
Earth (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-12
DAY 3 - APRIL 28, 2016 (PORTOVIEJO)
Team 1: A. Zekkos, K. Rollins, C. Wood, R. Luque, D. Alzamora, A. Caicedo.
Objective: Bridge behavior in Rocafuerte and Portoviejo, observations of lateral spreading and
measurements of Vs.
Route: GPS track is shown on Fig. 2.3.4.

Figure 2.3.4. Team 1, 4/28/16: GPS track to observe behavior of bridges in Rocafuerte and Portoviejo.
Base Image: Google Earth (©2016).

Team 2: S. Nikolaou, X. Vera, R. Gilsanz, V. Diaz, C. Arteta, G. Lyvers, G. Ponce.


Objective: Observe essential structures (hospitals, schools), Portoviejo ground zero, cemetery.
Team 3: R. Gilsanz, C. Arteta, V. Diaz, A. Caicedo, G. Ponce.
Objective: Jaramijo Port, Manta Airport (end of day).
Route: GPS tracks for Teams 2 and 3 are shown on Fig. 2.3.5.

Figure 2.3.5. Teams 2, 3, 4/28/16: GPS tracks for essential buildings. Base Image: G. Earth (©2016).
GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance
Report Version 1 2-13
DAY 4 - APRIL 29, 2016 (PEDERNALES)
Team 1: K. Rollins, C. Wood, G. Ponce, R. Luque, D. Alzamora, G. Lyvers
Objective: Evaluate damage on route to Pedernales including Bahia de Caraquez and Los Caras
Bridge (Hwy 15), La Fortuna Bridge and the towns of Jama and Pedernales. Team looked for
signs of liquefaction and general damage of these areas.
Route: GPS track is shown on Fig. 2.3.6 and surveys routes in Jama and Pedernales are mapped
on Figs. 2.3.7, 2.3.8.

Figure 2.3.6. Teams 1, 2, 4/29/16: GPS route to Pedernales, Base Image: Google Earth (©2016).

Figure 2.3.7. Team 1, 4/29/16: GPS survey route in Jama. Base Image: Google Earth ( ©2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-14
Figure 2.3.8. Team 1, 4/29/16: GPS survey route in Pedernales. Base Image: Google Earth ( ©2016).

Team 2: X. Vera, R. Gilsanz, V. Diaz, C. Arteta, A. Caicedo.


Objective: Evaluate damage on route to Pedernales including Bahia de Caraquez and Los Caras
Bridge (Hwy 15), Los Perales airport in San Vicente and the towns of Canoa and Pedernales.
Route: GPS track is shown on Fig. 2.3.6 and routes of Los Perales Airport in San Vicente,
Canoa, and Pedernales are mapped on Figs. 2.3.9, 2.3.10, and 2.3.11.

Figure 2.3.9. Team 2, 4/29/16: GPS survey route of Los Perales Airport in San Vicente. Base Image:
Google Earth (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-15
Figure 2.3.10. Team 2, 4/29/16: GPS survey route in Canoa. Base Image: Google Earth ( ©2016).

Figure 2.3.11. Team 2, 4/29/16: GPS survey route in Pedernales. Base Image: Google Earth ( ©2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-16
DAY 5 - APRIL 30, 2016 (FLYOVERS)

All team members travel via military helicopter generously provided by the Ecuadorian
Armed Forces in the paths mapped on Fig. 2.3.12. The team left from the Portoviejo Airforce
base and flew over coastal damaged area to observe landslides and liquefaction and the dams
Poza Honda, La Esperanza, and Chone. A second flyover was made to Pedernales flying over
Canoa and Jama. On the way back, the helicopter followed the coastal road south, circled Bahia
de Caraquez and landed back in Portoviejo. More details on the flyovers are provided on
Chapter 6. Following the helicopter observations, Team 1 visited Chone Dam.

Figure 2.3.12. GEER-ATC Team, 4/30/16: Helicopter paths Base Image: Google Earth ( ©2016).
Team 1: K. Rollins, C. Wood, G. Lyvers, A. Caicedo, D. Alzamora.
Objective: Chone Dam site visit and shear wave velocity Vs and HVSR measurements.
Route: GPS track for Canoa Dam visit is shown on Fig. 2.3.13.

Figure 2.3.13. Team 1, 4/30/16: GPS track of Vs surveys at Chone dam. Base Image: G. Earth (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-17
DAY 6 – MAY 1, 2016 (MANTA)

Team 1: S. Nikolaou, X. Vera, R. Gilsanz, V. Diaz.


Objective: Meeting with Ministry of Development and Housing (Ministra M. Duarte) to brief
them of general GEER-ATC observations and discuss recovery and rebuilding planning.
Team 2: K. Rollins, D. Alzamora.
Objective: Observations and settlement measurements in Manta ground zero.
Team 3: C. Wood, G. Lyvers.
Objective: Grid measurements of HVSR in Manta ground zero, following suggestions from the
structural team member’s observations. Data to be used to correlate structural damage.
Route: GPS track for Teams 2 and 3 are shown on Fig. 2.3.14.

Figure 2.3.14. Teams 2, 3, 5/1/16: GPS locations of Vs surveys at Manta ground zero. Base Image:
Google Earth (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 2-Ecuador and Reconnaissance


Report Version 1 2-18
Blank Page
Blank Page
CHAPTER 3
GEER-ATC
Reconnaissance Team

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
3 GEER - ATC Team
INTRODUCTION

The main Mw7.8 earthquake event hit Ecuador on April 16th, 2016, centered offshore of the
west coast of northern Ecuador with many aftershocks following it. Based on the US
Geological Survey (USGS), the main event was caused by shallow thrust faulting on or near
the plate boundary between the Nazca and S. America plates where the Nazca subducts beneath
the S. America plate at 61 mm/yr. Subduction along the Ecuador and the Peru-Chile trenches
farther south, has led to uplift of the Andes mountains and has produced some of the largest
known earthquakes, including the largest ever recorded M9.5 Chile earthquake in 1960.

The event drew the attention of the Geotechnical Extreme Events Reconnaissance (GEER)
Association, due to the several hundred casualties, tens of thousands homeless, and destruction
along the west coast, with evidence of severe ground motions and geotechnical failures. GEER
had activated a reconnaissance team with funding by the National Science Foundation (NSF)
when a second major Mw6.1 event occurred on April 20th. The GEER team was joined by
structural engineers funded by the Applied Technology Council (ATC). The US-based team
was on the ground from April 26th to May 2nd, joined by Ecuadorian counterpart partners.

The mission to Ecuador collected reconnaissance data in an organized, geo-tagged manner


that can help to investigate the causes of widespread geotechnical and infrastructure failures,
intense ground motions affected by the regional soils, liquefaction, soil-structure interaction
effects, and structural and nonstructural damage. Information on local practice, rapid
assessment procedures, and community preparedness and response was observed. The team
documented, in addition to failures, cases of success in an effort to understand resilient
behavior under extreme events.

The 2016 Ecuador GEER-ATC mission was a collaboration of an international multi-


disciplinary team of more than 40 people affiliated with institutions, universities, and private
firms. This Chapter acknowledges all involved parties and explains their role, with contributing
authors in the report denoted by (CA) and editors by (E).

NOTE: The editors of this report, Dr. Sissy Nikolaou, Dr. Xavier Vera-Grunauer, Mr. Ramon
Gilsanz, have tried to acknowledge all individuals and organizations that worked with or assisted our
team, both in Ecuador and US, volunteered time in providing, collecting, and organizing information,
or co-authoring. We will correct any (unintentional) omission in the next version of this report.

GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team


Report Version 1 3-1
USA TEAM

The GEER-Ecuador team was co-led by Dr. Sissy Nikolaou of WSP | Parsons Brinckerhoff
(WSP|PB) from the United States (US) and Dr. Xavier Vera-Grunauer of the Universidad
Católica de Santiago de Guayaquil and of the Geoestudios firm of Ecuador (EC). The NSF-
funded GEER members that travelled to Ecuador were geotechnical engineers Mr. Daniel
Alzamora (Federal Highway Administration), Prof. Adda Athanasopoulos-Zekkos (Univ. of
Michigan), Ms. Gabriela Martinez Lyvers (US Army Corp of Engineers), Prof. Kyle Rollins
(Brigham Young Univ.), and Prof. Clint Wood (Univ. of Arkansas). The US team members
funded by ATC were structural engineers Mr. Ramon Gilsanz of Gilsanz Murray Steficek
(GMS) and Ms. Virginia Diaz (GMS), who was the ATC recorder for the report preparation.

The GEER-ATC group was teamed with other US team members, already in Ecuador,
funded individually or by their institutions, including Prof. Eduardo Miranda (Stanford Univ.),
Mr. Enrique Morales (Univ. at Buffalo & L. Colonel of EC Army Corps of Engineers), and
Mr. Roberto Luque (Univ. of California at Berkeley). During reconnaissance in Ecuador, Mr.
Guillermo Diaz-Fanas (WSP|PB) served as GEER-Ecuador recorder for the information
gathered and the USA contact person along with Mr. Pablo Lopez (PE) who facilitated
communications while aboard. The GEER-ATC USA team was joined by Ecuadorian counter-
partners from the government, army, universities, and private engineering firms (presented in
Ecuador Team section) and by Colombian partner from Universidad Norte, Prof. Carlos Arteta.
The Ecuador-USA GEER-ATC team met daily to discuss findings and plan the next day.

While in Ecuador, the US team members had the unique opportunity to interact with local
government officials and discuss reconnaissance findings, and rebuild needs. Several meetings
were held in Guayaquil and Quito with the Minister of Urban Development & Housing
(MIDUVI) Ms. Maria de los Angeles Duarte and her staff. The MIDUVI was instrumental by
arranging access to sites and providing offices, maps of damage, and post-reconnaissance
information. Our team also met with representatives from other Ministries, administrators of
hospitals, schools, the 911 Emergency Center, and the Ecuadorian Society of Civil Engineers.

The Ecuadorian Army Corps of Engineers (ECACE) provided invaluable support by


arranging helicopter flyovers, providing data of rapid assessment and design documentation
for the case studies that we observed. The ECACE leadership, including Commander General
Pedro Mosquera, Colonel Jose Ramos, Colonel Xavier Riofrio, and Lieutenant Colonel
Enrique Morales, was in constant contact with our team and facilitated logistics and data needs.

GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team


Report Version 1 3-2
USA Funding Agencies
Financial support for GEER was provided by the National Science Foundation (NSF). The
Applied Technology Council (ATC) funded the participation of two structural engineers.
Collaborating Universities
Several US universities supported the GEER-ATC mission by providing resources of data
and equipment, and providing volunteer time of their staff and students (see also Chapter 4).
BERKELEY Jonathan D. Bray, PhD, PE, F. ASCE, NAE, GEER Chair
Faculty Chair in Earthquake Engineering Excellence
Civil and Environmental Engineering
Roberto LuqueCA, PhD Candidate
Civil and Environmental Engineering
Tadahiro Kishida CA, PhD
Assistant Project Scientist and Post-doctoral researcher
Pacific Earthquake Engineering Research (PEER) Center
BYU Kyle RollinsCA, PhD, GEER Member
Professor, Brigham Young University
Civil and Environmental Engineering
COLUMBIA Jenny (Evgenia) Sideri, Former PhD Candidate
Civil and Environmental Engineering
Athina Spyridaki, Former PhD Candidate
Civil and Environmental Engineering
CORNELL Thomas D. O'RourkeCA, PhD, Dist. M.ASCE, NAE, FREng
Thomas R. Briggs Professor of Engineering, GEER Advisory Panel
School of Civil and Environmental Engineering
Bernardo CasaresCA, Undergraduate Student
School of Civil and Environmental Engineering
GATECH J. David Frost, PhD, PE, F. ASCE, GEER Co-Chair
Elizabeth & Bill Higginbotham Professor
Civil and Environmental Engineering
ILLINOIS UC Youssef Hashash, PhD, PE, F.ASCE, GEER Steering Com.
W.J. & E.F. Hall Endowed Professor, J. Burkitt Webb Faculty Scholar
Civil & Environmental Engineering, Urbana-Champaign
RPI Nonika AntonakiCA, Former PhD Candidate
Civil and Environmental Engineering
Panagiota KokkaliCA, Former PhD Candidate
Civil and Environmental Engineering
STANFORD Eduardo MirandaCA, PhD
Professor, Associate Department Chair
Civil and Environmental Engineering
Andrés Acosta, PhD candidate
Civil and Environmental Engineering
Luis Ceferino, PhD candidate
Civil and Environmental Engineering

GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team


Report Version 1 3-3
UARK Clinton WoodCA, PhD, GEER Member
Assistant Professor
Civil Engineering, Univ. of Arkansas
UNCO Carlos ArtetaCA, PhD
Professor, Civil Engineering Dept.
Universidad Norte Colombia, Barranquilla, Colombia
UBUFFALO Andrew Whittaker, PhD, PE, SE, F.SEI, F.ASCE
Professor and Chair
Civil, Structural and Environmental Engineering, University at Buffalo
Michael C. Constantinou, PhD, SUNY Distinguished Professor
Samuel Capen Professor
Civil, Structural and Environmental Engineering
Enrique MoralesCA, PhD Candidate
Civil, Structural and Environmental Engineering
Lieutenant Colonel, Army Corps of Engineers of Ecuador
Jerome S. O'ConnorCA, PE
Adjunct Professor
Executive Director of Institute of Bridge Engineering
Andreas StavridisCA, PhD
Assistant Professor
Civil, Structural and Environmental Engineering
Jan Diaz, Graduate Student
Civil, Structural and Environmental Engineering
UMICHIGAN Adda Athanasopoulos-ZekkosCA, PhD, GEER Member
Associate Professor
Civil and Environmental Engineering
Oscar GonzálezCA, Graduate Student
Civil and Environmental Engineering
UTEXAS Ellen Rathje, PhD, PE, GEER Co-Chair
Warren S. Bellows Centennial Professor
Civil, Architectural, Environmental Engineering, U-TX Austin
VTECH Roberto León, PhD, F.ACI, ATC Board Member
David H. Burrows Professor
Charles E. Via, Jr. Department of Civil & Environmental Engineering
USA Agencies, Organizations, and Individual Contributions
Organizations and individuals supported the team with resources, data, report preparation:
USACE United States Army Corps of Engineers
Gabriela M. LyversCA, PE
Structural Engineer
FHWA US Federal Highway Administration
Daniel AlzamoraCA, PE
Geotechnical Engineer
ACI American Concrete Institute
Esteban Anzola, PE
ACI 314 Committee Member (Spanish version for Latin America)

GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team


Report Version 1 3-4
ATC Applied Technology Council
Ayse Hortacsu, PE
Director of Projects
Veronica Cedillos
Associate Director of Projects
Roberto León, PhD, F.ACI
Board Member, Virginia Tech Professor of Civil Engineering
Jon Heintz, PE, SE
Executive Director
Victoria Arbitrio, PE, SECB, F.SEI
President 2016-2017
EERI Earthquake Engineering Research Institute
Marjorie Greene
Interim LFE Coordinator
FEMA Federal Emergency Management Agency
Eric Letvin
Deputy Assistant Administrator, Mitigation
INDIVIDUALS Pablo LópezCA, PE, GEER Volunteer
Foundation and Structural Engineer
Said MaaloufCA. GEER Volunteer
MCEER Multidisciplinary Center for Earthquake Engineering Research
Andrew Whittaker, PhD, PE, SE, F.SEI, F.ASCE
Director
NYCOEM NYC Office of Emergency Management
Heather Roiter
Director, Hazard Mitigation
Melissa Umberger
Project Manager, Hazard Mitigation Project Manager
Cynthia Barton
Program Manager, Housing Recovery Program Manager
NYCHRO NYC Mayor’s Housing Recovery Office
Susan Rosenstadt
Deputy Director of Planning, Design & Technical Services
PEER Pacific Earthquake Engineering Research Center
Tadahiro Kishida CA, PhD
Post-doctoral Researcher
UGC Urban Green Council
Cecil Scheib, PE, CEM, LEED AP, GPRO O&M
Chief Program Officer
USGS United States Geologic Survey
David Wald, PhD
Seismologist

GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team


Report Version 1 3-5
WH National Security Council, The White House
Judge Alice C. Hill (ret.)
Special Assistant to President Obama, Sr. Director for Resilience Policy
David V. Adams
Director for Health Security and Climate Resilience Policy
USA Engineering Private Firms
The private firms that provided staff time, resources, report/presentations preparation are:
GMS Gilsanz, Murray, Steficek, LLP, New York, New Jersey, Los Angeles
Ramon GilsanzE,CA, PE, SE, F.SEI, F.ASCE
Founding Partner, Structural Engineer, ATC Representative
Virginia DiazCA, PE
Senior Structural Engineer, ATC Recorder, GEER-ATC Ecuador
Laura HernandezCA
Structural Engineer
Connie Yang, Staff member
WSP|PB WSP | Parsons Brinckerhoff, New York, Miami, Seattle, Mexico
Sissy NikolaouE,CA, PhD, PE, F.ASCE, GEER Advisory Committee
Principal, Multihazards and Geotechnical Engineering, NY
Co-leader, GEER-ATC Ecuador Mission
Frank Pepe Jr., PE
USA Director of Geotechnical & Tunneling Technical Excellence Center
Dale Moeller, PE
Vice President, Geotechnical & Tunneling Technical Excellence Center
Guillermo Diaz-FanasCA
Geotechnical Engineer, G&T Tech. Center, GEER-Ecuador Recorder
Nonika AntonakiCA, PhD
Geotechnical Engineer, NYC G&T Technical Excellence Center
Panagiota KokkaliCA, PhD
Geotechnical Engineer, NYC G&T Technical Excellence Center
Patrick BassalCA, PE
Geotechnical Engineer, NYC G&T Technical Excellence Center
Efthymios (Themis) VaxevanisCA
Senior Virtual Design & Construction Engineer, NYC
Jay MezherCA, AIA
Director, Virtual Design & Construction, Seattle
Jeff Smilow, PE, F.ASCE
Executive Vice President of Building Structures, NYC
Ahmad Rahimian, PhD, PE, SE, F.ASCE
USA Director of Building Structures, NYC
Esteban Anzola, PE
Vice President, Building Structures, Miami, ACI-318 Committee Member
Gerardo Aguilar, PhD
Building Structures, NYC
Rodolfo Valles Mattox, PhD
Director General, México
GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team
Report Version 1 3-6
ECUADOR TEAM

Ecuadorian universities and institutions involved with earthquake engineering volunteered


in the GEER-ATC mission by providing available information and data collected by them prior
to the US team arrival, participating in reconnaissance and acting as external reviewers. Some
of the Ecuadorian team members had already performed reconnaissance immediately after the
main event and offered valuable input. Local practicing geotechnical and structural engineers
volunteered their time in the field, provided resources and performed testing, collected data,
and participated in the report preparation. All parties are acknowledged below as academic or
government institutions, or individuals with their affiliations.

Ecuadorian Public Agencies


MIDUVI Ministry of Urban Development and Housing of Ecuador
Maria de los Angeles Duarte, Architect
Minister of Housing
Gonzalo Vela, Logistic Leader of Corps of Engineers
COE Technical Team, PNUD, United Nations
Carolina Salcedo, Regional Coordinator
COE Technical Team
Veronica Bravo, Civil Engineer
Sub director of Ministry of Housing
MTOP Ministry of Transportation & Public Works of Ecuador
Walter Solis, Civil Engineer
Minister of Transportation & Public Works
SA Carola Gordillo, Technical Advisor
Water Secretariat, EC
C. Bernal, Civil Engineer
Water Secretariat, Ecuador
Ecuadorian Army Forces
ECACE General Pedro Mosquera
Commander of EC Army Corps of Engineers
Colonel Jose Ramos
Colonel of EC Army Corps of Engineers
Colonel Xavier Riofrio
Logistic Leader of EC Army Corps of Engineers
Lieutenant Colonel Enrique Morales MoncayoCA
ECACE and PhD Candidate, University at Buffalo, USA
Universities and Individuals
ESPE Theofilos ToulkeridisCA, PhD
Professor and Researcher of Geology, Geochemistry, Volcanology
Escuela Politecnica del Ejercito, & Univ. de las Fuerzas Armadas, EC

GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team


Report Version 1 3-7
ESPOL Pedro Rojas-Cruz, PhD
Professor, Civil Engineering Department
IG-EPN Hugo YepesCA, PhD
Professor and Researcher of Geophysics
Escuela Politecnica Nacional, Instituto Geofísico, EC
Alexandra AlvaradoCA, PhD
Head of Seismological Branch of National Geophysical Institute
Escuela Politecnica Nacional, Instituto Geofísico, EC
SFQUITOU Fabricio YepezCA, PhD
Vice-Dean of Engineering School and Professor
Civil Engineering, IIFIUC
Universidad San Francisco de Quito, EC
UCSG Adolfo CaicedoCA
Geotechnical Engineer Researcher, Civil Engineering Dept.
Universidad Católica de Santiago de Guayaquil, Ecuador (UCSG)
Claudio Luque
Professor, Civil Engineering Dept.
Walter Mera, PhD
Vice President of UCSG
Guillermo Ponce
Professor, Civil Engineering Dept.
Francisco RipaldaCA
Professor, Civil Engineering Dept.
Xavier Vera-GrunauerE,CA, PhD, D.GE
Head of the Engineering Research Institute, IIFIUC
UNCO Carlos ArtetaCA, PhD
Professor, Civil Engineering Dept.
Universidad Norte Colombia, Barranquilla, Colombia
INDIVIDUALS Karina Galvez, GEER Volunteer
TV anchor at UCSG Television, Poet, Photographer

Infrastructure Data Contributors


Bahia Bridge CPR Asociados
Adolfo Caicedo
Nylic
Claudio Luque
Mejia Bridge ECACE, Ecuadorian Army Corps of Engineers
Lieutenant Colonel Enrique Morales Moncayo
Nylic
Constructora Verdu
AET
Jaime Pesantes
Geoestudios, SA
Xavier Vera-Grunauer, PhD, D.GE

GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team


Report Version 1 3-8
Los Caras Br. ECACE, Ecuadorian Army Corps of Engineers
Lieutenant Colonel Enrique Morales Moncayo
Los Chacras CPR Asociados
Rio Chico Br. CPR Asociados
Adolfo Caicedo
Chone Dam GEER-ATC Team
Daniel Alzamora
Manta Port UCSG, Universidad Católica de Santiago de Guayaquil
Oswaldo Ripalda
Geoestudios, SA
Xavier Vera-Grunauer, PhD, D.GE
Jaramijo Port UCSG, Universidad Católica de Santiago de Guayaquil
Oswaldo Ripalda
Facilities ECACE, Ecuadorian Army Corps of Engineers
Bahia Hospital
Chone Hospital
IESS (Ecuadorian Institute of Social Security) Manta Hospital
Manta Rodríguez Zambrano Hospital
Solca Hospital
Ceibo Dorado Hotel
Supreme Court of Justice Building
Ministry of Public Health
Manta Rodríguez Zambrano Hospital
María Beatrix Santos Velez

Engineering Firms
GEOESTUDIOS Xavier Vera-GranauerE,CA, Prof. Civil Engineer, PhD, D.GE
Founder and Owner
Noemí Villagrán LeónCA
Assistant
Ignacio OchoaCA
Civil Engineer
Danilo DavilaCA
Technical Assistant
DRONEANDO Luis Fernando RodríguezCA
Owner
CVA Consultora Vera & Asociados
Juan Vera, Professional Civil Engineer
Technical Director
Francisco Vera González, Professional Civil Engineer
CEO, External Contributor to the Report

GEER-ATC Muisne, Ecuador, Earthquake 3-Reconnaissance Team


Report Version 1 3-9
Blank Page
CHAPTER 4
Seismotectonics
Seismic History

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
4.1 Seismotectonic Setting
OVERVIEW

Despite being one of the smallest countries in the world, Ecuador is one of the few, if not
the only place, that exhibits all forms of earth movements represented by the three different
types of plate boundaries. This geodynamic setting presents the opportunity to study and
understand the evolution of huge mountain ranges such as the Cordillera Real and the volcanic
features represented by the continental volcanoes and the Galapagos Islands as a result of
subduction and hot spot volcanism, respectively.

There are two divergent plate boundaries in the Pacific Ocean: the East Pacific Rise and
Galapagos Spreading Center, in which new oceanic crust is continuously being generated. In
these areas, three distinctive oceanic crusts originate (Fig. 4.1.1): the so-called Pacific, Cocos
and Nazca oceanic plates (on a triple junction around the Galapagos microplate), which move
away from each other at the same rate as fingernails grow. While these plates move away from
each other, the Nazca Oceanic plate moves almost eastwards and collides along a convergent
zone with the coast of the South American continent.

The presence of two major bathymetric elements on the Nazca plate complicates its
penetration beneath the continent (Yepes et al., 2016): (a) the Carnegie ridge, a ~200 km wide,
2 km high volcanic ridge originated at the Galapagos hot spot; and (b) a ~500 m high oceanic
floor step characterizing both sides across the Grijalva ridge (Fig. 4.1.2). The Grijalva ridge is
the remaining scar and final evidence of the Farallon plate’s rifting process at the beginning of
the Miocene (Hey, 1977; Lonsdale, 2005) and is named the Grijalva Rifted Margin (GRM).
There is a density contrast due to up to 9 million years (Ma) age difference at both sides of the
GRM, responsible for the bathymetric step between the Nazca and Farallon plates.

Both the Carnegie ridge and GRM have been entering the subduction zone from at least 3-
6 Ma ago and are thought to have penetrated more than 300 km under the continent (Gutscher
et al., 1999; Michaud et al., 2009). This implies that actually two plates with different densities
bonded along a weak zone (GRM) and are being subducted underneath continental Ecuador.
Due to the acute convexity of the continental margin (Fig. 4.1.2), the oblique convergence
resulting from the margin’s shape and the high interseismic coupling along the subduction
interface, two different continental slivers are driving away from each other (Nocquet et al.,
2014; Yepes et al., 2016).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-1
Report Version 1
GEER-ATC Muisne, Ecuador, Earthquake
Figure 4.1.1. The geodynamic setting of Ecuador and associated plates, microplate and volcanic ridges (Oceanic Plates: Pacific, Nazca, Cocos; Continental
Plates: Caribbean and South American; Oceanic Microplate: Galapagos, Ridges: Carnegie, Cocos (after Toulkeridis, 2013).

4-Seismotectonics and Historic Seismicity


4-2
The North Andean Sliver (NAS) is moving toward the NNE at 9 mm/yr along localized
right-lateral, strike-slip or reverse faults that continue in Colombia as the Algeciras fault (the
CCPP system, red thick line in Fig. 4.1.2). To the south, the Inca Sliver is moving toward the
SSE at ~5 mm/yr along the proposed limit in the eastern Peruvian sub-Andean Belt (dashed
red line in Fig. 4.1.2, Nocquet et al., 2014). The CCPP fault system is responsible for major
crustal earthquakes in continental Ecuador.

Figure 4.1.2. Geodynamic setting of Ecuador with main cities shown with white letters: Pa=Pasto;
Q=Quito; R=Riobamba; B=Bahía; G=Guayaquil; Cu=Cuenca; Ta=Talara; Bu=Buenaventura. Base
map modified from GeoMapApp (geomapapp.org).

The South American continent itself is composed of two continental plates, the Caribbean
and South American plates, neither of which move toward nor away from each other. They
touch each other and move in opposite directions along the third type of plate boundary. This
transformation or strike-slip fault, which extends from the Gulf of Guayaquil through
Venezuela, is called the Guayaquil-Caracas mega-fault or shear and is responsible for
destructive earthquakes along and aside the fault.

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-3
The sea floor on the Nazca plate has variable depth, as the 11 major Fracture Zones (FZ)
of the plate separate the crust into sections of different ages as seen on Fig. 4.1.3 (Müller et al.,
1997) along with Ridges (R) and locations of six subduction earthquakes of Mw ~ 8.3 and larger
since 1900 (Carena, S., 2011) with epicenter near a major FZ or R on the subducting plate.

Figure 4.1.3. Ocean floor ages along South America’s Nazca plate with major Fracture Zones (FZ) and
Ridges (R), and locations of 6 subduction earthquakes of Mw greater than 8.3 since 1900 with epicenter
(in white circle) near a FZ or R and focal plane shown where available. JFR = Juan Fernandez R, IQR
= Iquique R, PA = Pacific, NZ = Nazca, SA = South America (Carena, S., 2011 and Müller et al., 1997).

Future earthquakes in Ecuador are anticipated along the Guayaquil-Caracas mega-fault,


and also as result of the collision and subsequent subduction between the Nazca oceanic and
Caribbean/South American continental plates. Along the coast of Ecuador, one can find
different pieces of accreted microplates or terranes that are remains of the upper part of
previous islands of the Galapagos forming a submarine chain of extinct volcanoes (seamounts)
called the Carnegie Ridge.

Regional tectonic characteristics of the Galapagos hot spot, continental Ecuador and its
volcanoes, and coastal Ecuador and its tsunamis will be described in detail in the next sections.

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-4
GALAPAGOS HOT SPOT

The main Galapagos islands appear above the Nazca plate around 250 km south of the EW
trending Galapagos spreading center and some 1,000 km west of the Ecuadorian mainland (Fig.
4.1.4). With an area less than 45,000 km2 , the islands are one of the most volcanically active
regions in the world and are a product of an apparent mantle plume beneath the region, also
called “hot spot.” Mantle plumes represent columns of hot melted rock that rise from the core-
mantle boundary some 2,900 km deep in the earth. Due to its upward motion, the mantle plume
pushes the overlying lithosphere upward creating an anomalously shallow region around the
islands called the Galapagos platform which may represent a basalt flood. The hot spot remains
fixed relative to the lithospheric plate moving above, but when it crosses the lithosphere it
supplies newly formed volcanoes with magma.

Figure 4.1.4. The Galapagos spreading center is an EW ridge rising above the seafloor with shallowest
depth ~1.6 km north of the Islands. Areas deeper than 3.5 km are shown in blue (Haymon et al., 1993).

At the Galapagos hot spot, this process of magma supply has existed for more than 90 Ma,
while the lithospheric plate has moved thousands of km, carrying generated volcanoes away,
and making them eventually extinct. Two aseismic volcanic ridges, the NE moving Cocos
Ridge and the eastern moving Carnegie Ridge and seamounts on the Cocos and Nazca plates,
are consistent with the hot spot theory. These extinct volcanic ridges are due to cooling-
contraction reactions of magma, which slowly sank below the sea surface due to lack of magma
supply, lithospheric movement and erosional processes. Hence, it appears that the volcanism
age increases eastward as the Nazca plate moves away from the hot spot.

Over time, the submarine volcanic ridges and various microplates amass at the South
American continent. Western islands just above the Galapagos hot spot have the morphology
of large shield volcanoes with deep calderas, while the eastern ones are small shield volcanoes

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-5
with gentle slopes and almost without calderas. An exception are the northern islands of Wolf
and Darwin, considered to be result of the interaction of the hot spot and the spreading center.

CONTINENTAL ECUADOR AND VOLCANOES

The northern Andes in Ecuador are part of the 7,000-km long active continental margin
with volcanic sequences of Mesozoic and Cenozoic ages (Ramos, 2009). The volcanic arc
appears in a NNE-SSW strike, and is the result of the perpendicular collision between the
Miocene oceanic Nazca plate and subducted with a slightly oblique angle to the southern
American continental lithosphere, which is itself made of the Caribbean and South American
continental plates (MéGard, 1987; Colmenares & Zoback, 2003; Dumont et al., 2005; Egbue
& Kellogg, 2010; Toulkeridis, 2013). The Nazca plate incorporates the aseismic Carnegie
Ridge, and is formed by the Nazca plate moving ESE over the Galapagos hot spot (Johnson &
Lowrie, 1972; Freymuller et al., 1993; Werner et al., 2003; Toulkeridis, 2011). Fig. 4.1.5
presents the Ecuador subduction zone and associated volcanic chains (Toulkeridis, 2013).

Figure 4.1.5. Subduction zone of Ecuador and its volcanic chains (after Toulkeridis, 2013).

The up to 250 volcanoes on the volcanic front, main, back and rear arc are within 4 zones:
the Western, Interandean, Eastern and Subandean cordilleras (Toulkeridis, 2013) and are all
part of the Andean northern volcanic zone (Barberi et al., 1988; Bryant et al., 2006; Hall et al.,
2008). Nearly 20 volcanoes were active in the last century, 5 erupting in the past 17 years (Fig.
4.1.6) and listed below.

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-6
i. Sangay with constant activity (with higher, registered intensity in 2004-11, 2013, 2016)
ii. Guagua Pichincha (1999-2001, 2009)
iii. Reventador (2002-2016 with major interruptions)
iv. Tungurahua (1999-2016 with less interruptions); and
v. Cotopaxi (2015).

Figure 4.1.6. Left: Guagua Pichincha (0°10'14.88"S, 78°36'45.36"W), 10/7/99; U. right: Tungurahua,
2002 (1°28'12.71"S, 78°26'41.28"W); Center: Sangay, 2001 (2°00'17.99"S, 78°20'26.88"W); L. right:
Reventador, 2002 (0°04'39.00"S, 77°39'20.87"W). Photos by A. Speck & Anonymous (Reventador).

Figure 4.1.7. Volcanic Explosivity Index (VEI) scale and ejecta volume (volcanoes.usgs.gov).

Details are provided in Monzier et al. (1999); Garcia-Aristizabal et al. (2007); Toulkeridis
et al. (2007); Arellano, et al. (2008); Barrancos et al. (2008); Carn et al. (2008); Hall et al.
GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity
Report Version 1 4-7
(2008); Ridolfi et al. (2008); Robin et al. (2008); Lees et al. (2008); Steffke et al. (2010);
McCormick et al. (2014); Toulkeridis et al. (2016); Vaca et al. (2016a,b,c,d). Table 4.1.1
presents continental Ecuador volcanoes with their activity and max Volcanic Explosivity Index
(VEI), which is a relative explosiveness measure based on volume of products, eruption cloud
height, and qualitative observations (Newhall & Self, 1982). The VEI log scale shown on Fig.
4.1.7 is open-ended with each interval of VEI greater than 2 representing 10-fold ejecta
increase. Non-explosive eruptions (< 10,000 m3 of tephra) have a VEI of 0; and a VEI of 8 is
a mega-colossal eruption with 1,012 m3 of tephra and cloud height of over 20 km.

Table 4.1.1. Active volcanoes of continental Ecuador and max VEI (Smithsonian Institute’s Global
Volcanism Program, volcano.si.edu/world/region.cfm?rnum=1502, and Toulkeridis, 2013).

max Volcanic Last Known active phases


Name
Explosivity Index (VEI) Erruption in past 20,000 years
Chacana 0 1773 4

Antisana 2 1802 3

Cerro Negro de Mayasquer 2 volcanic unrest 1?

Chiles 2 volcanic unrest 1?

Aliso 2 (?) 2450 BC (?) 1?

Chachimbiro 2 3200 ± 20 BC 1

Imbabura 2 5500 ± 500 BC 1

Sangay 3 2016 3

Cayambe 4 1785 22

Chimborazo 4 550 ± 150 6

Reventador 4 2016 28

Niñahuilca 5 320 ± 16 BC 4

Guagua Pichincha 5 2009 43

Tungurahua 5 2016 32

Cotopaxi 5-6 2015 84

Sumaco 3-6 1933 (?) 3

Cuicocha 5-6 6650 BC (?) 4

Soche 5-6 1797? 1

Quilotoa 6 6 6

Pululahua 6 290 BC 4

Chalupas 7 18,000 BC (?) 1?

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-8
COASTAL ECUADOR AND TSUNAMIS

Due to its geodynamic setting, the Ecuadorian continental platform, similar to other
countries along the Pacific rim, is exposed to tsunamis (Gusiakov, 2005). The active
continental margin and subduction zone between the oceanic Nazca and continental South
American and Caribbean plates (Fig. 4.1.1), both separated by the Guayaquil-Caracas mega
shear (Kellogg & Vega, 1995; Gutscher et al., 1999; Egbue & Kellog, 2010), generate tsunamis
of tectonic or submarine landslide origin (Shepperd & Moberly, 1981; Pontoise & Monfret,
2004; Ratzov et al., 2007, 10; Ioualalen et al., 2011; Pararas-Carayannis, 2012). The Galápagos
volcanism may be another origin of tsunamis (Toulkeridis, 2011). The Galápagos hot spot has
produced many voluminous shield-volcanoes which are most inactive now due to the ESE-
movement of the overlying Nazca plate (Holden & Dietz 1972; Toulkeridis, 2011).

The main Galápagos Islands are located south of the E-W-trending Galápagos Spreading
Center, east of the NS-trending East Pacific Rise and some 1,000 km west of the mainland.
Due to volcanic activity and plate drifting, two aseismic ridges were created: the Cocos moving
NE above the Cocos plate and the Carnegie moving east of the Nazca plate (Harpp et al., 2003).
Aseismic ridges like the Carnegie Ridge of Fig. 4.1.8 become an obstacle in the oblique
subduction process and may generate a potential valve of tsunamis along the Ecuadorian coast
within the subduction zone. The Carnegie Ridge collides towards the continental margin with
a velocity of as low as 5 cm/yr between 1°N and 2°S in latitude (Pilger, 1983). Further local
tsunamis may be generated by seismotectonic extensional processes along the Ecuadorian coast
like in the Gulf of Guayaquil (Pararas-Carayannis, 2012).

Figure 4.1.8. Carnegie Ridge in 3-D imaging.

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-9
In the past two centuries, the Ecuadorian shoreline has experienced a dozen impacts of
tsunamis from local origins of various intensities with one being up to M w8 in 1906 (Rudolph
& Szirtes, 1911; Kelleher, 1972; Beck & Ruff, 1984; Kanamori & McNally, 1982; Swenson
& Beck, 1996; Pararas-Carayannis, 2012), while evidence of paleo-tsunami deposits is scarce
(Chunga & Toulkeridis, 2014). Other prominent examples of tsunamis along the Ecuador–
Colombia subduction zone include tsunamis in 1942 (Mw7.8), 1958 (Mw7.7) and 1979 (Mw8.2)
within the 600-km long rupture area of the great 1906 event (Collot et al., 2004).

The epicenters of last century’s tsunamis are presented on Fig. 4.1.9. While the 1906 event
took the life of up to 1,500 people in Ecuador and Colombia, but with unknown financial
impacts, the 1979 tsunami killed at least 807 people in Colombia and destroyed approximately
10,000 homes, electric power and telephone lines (Pararas-Carayannis, 1980) especially at the
port of Tumaco. A 200-km section of the Ecuador-Colombia coast subsided about 1.5 m and
uplift occurred offshore on the continental slope (Herd et al., 1981; Kanamori & McNally,
1982). A historic photo of the 1979 destruction is shown on Fig. 4.1.10.

Figure 4.1.9. Tsunamis epicenters as red points during the last century (Toulkeridis, 2013).

Marine quake studies suggest a major earthquake in this margin region, based on
substantial strain accumulation (Pararas-Carayannis, 2012), is highly likely to occur. A
potential similar event to the 1906 tsunami, may be even more destructive, if it happens near
high tide. Potential losses along the Ecuadorian coast are presently higher than in the past, due
to a high population density resulting from the growth of the fishing and tourism industries.

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-10
Figure 4.1.10. Photo of damage after the 1979 tsunami in Colombia (credit: El Pais, Colombia).

4.2 Historic Seismicity


OVERVIEW

Ecuador has a long history of earthquakes, both of volcanic and tectonic (interplate and
intraplate) nature. According to Ecuador’s Geophysical Institute, there have been at least 37
earthquakes of magnitude 7 or higher since 1541, when written records by the Spanish were
first maintained. Figure 4.2.1 shows maps of events with reported intensity of VI or more
(Egred, 2009). Presently the country is densely instrumented to record seismic events.

UNIFIED EARTHQUAKE CATALOGUE DATA

According to Beauval et al. (2013), the first information for coastal events was available at
the end of the 19th century. The same authors stated that the seismicity distribution reflects the
two major tectonic features in Ecuador: subduction of the Nazca plate and the active crustal
deformation occurring in the Andean Cordillera. Figure 4.2.2 presents events with minimum
magnitude of 4 from the unified earthquake catalog (1587–2009).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-11
Figure 4.2.1. Intensity database for Ecuador (Egred, 2009), containing events with at least one intensity
VI reported (from Beauval et al. 2013).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-12
Figure 4.2.2. Epicenters from the unified earthquake catalog 1587–2009, integrating instrumental and
historical earthquakes, displaying magnitude 4 and above (after Beauval et al., 2013).

2016 EARTHQUAKE RUPTURE AND SEISMIC GAP

Historic seismic events related to the 2016 earthquake rupture zone are shown on Fig. 4.2.3
(Beauval et al., 2013). The background instrumental seismicity between 1900 and 2009 with
Mw < 7 is plotted as open circles and the M w ≥ 7 as yellow stars. Years and magnitudes for the
largest earthquakes are in the upper left-hand corner. Open-circle sizes are proportional to the
seismic moment, Mo , of each event.

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-13
The 2016 earthquake is the purple star and aftershocks and post-earthquake seismicity
outside the rupture area are colored as shown in the lower right-hand corner. The pink patch is
the aftershock zone for the 1998 M w7.1 earthquake (BEZ). The Ecuador - S. Colombia
subduction interface is very active between Lat. 0.5ºS and ~500 km north. The seismic moment
release rate along this segment is of the highest in the world.

Figure 4.2.3. Current and historical seismicity related to the 2016 earthquake rupture zone (Beauval et
al., 2013). Main cities: Q=Quito; M=Manta; P=Portoviejo; G=Guayaquil.

As illustrated on Fig. 4.2.4, megathrust earthquakes occurred within the 500-km long
rupture area of the 1906 earthquake in 1942, 1958, 1979 and 1998. Chlieh et al. (2014)
discovered that the asperity located between the 1942 and 1958 ruptures remained unbroken

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-14
since the 1906 event and stated that a moment deficit equivalent to a M w between 7.5 and 7.7
event had accumulated and should be considered in the next decades, including a potential
tsunami. The seismic gap remained unbroken until the recent 2016 events (Fig. 4.2.4).

The zones of the 1942 and 2016 aftershocks bound, but do not overlap, the Bahía
Earthquake Zone (BEZ; Figs. 4.2.3 & 4.2.5). The BEZ may have acted as a barrier for the
southern propagation of the 1942 and 2016 ruptures and maybe for the 1906 mega-event as
well. The 200-km wide Carnegie ridge constitutes a barrier for large earthquakes to propagate
southwards, similar to other subducting bathymetric features along the South American
subduction zone (Bilek, 2010).

2016 (7.8)

Figure 4.2.4. Image from minimum ISC model by Chlieh et al. (2014). The 2016 megathrust earthquake
almost closes the gap that remained unbroken since the 1906 rupture.

Figure 4.2.5. Ecuador historic seismicity (modified from Swenson and Beck, 1996).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-15
DETAILS ON HISTORIC EVENTS

Historic events with M > 6 dated 1556 to 2016 from the NOOA earthquake database are
summarized on Tables 4.2.1 & 4.2.2. Details for significant events are presented in this section.

Table 4.2.1. Historic events with M > 6 from 1556 to 2016 (NOOA earthquake database, 2016).

DATE TIME LOCATION PARAMETERS DEATHS DAMAGE COLLAPSE


1 3
Year Mo Dy UTC Name LAT LON D (km) M MMI No. Dscr. $M Dscr. No. Dscr.4

1556 10 17 Ecuador -0.200 -78.600 7.6

1575 9 8 Ecuador -0.200 -78.600 7.8 3

1587 8 30 Quito -0.220 -78.500 9 3

1590 8 7 Quito -0.500 -78.200 2

1627 6 26 11:00 Quito -0.200 -78.500 8 2

1640 2 Ecuador -1.700 -78.600 5,000 4 3

1641 1 10 Ecuador -1.500 -78.500 7.6

1645 2 19 Ecuador -1.700 -78.600 7.5 3

1660 10 27 14:00 Ecuador -0.200 -78.500 7.7

1662 1 1 Ecuador -0.200 -78.400 2

1674 8 29 Chimborazo -1.700 -79.000 7.7 10

1687 11 22 Ecuador -1.300 -78.600 7.3 2

1689 3 Tixan -2.200 -78.900 5 6.8 10

1698 6 20 6:00 Tunguruhua -1.200 -78.700 7.7 10 3

1698 7 19 Ambato -1.300 -78.700 1,000 3 3

1736 12 5 Ecuador -0.900 -78.600 2

1755 4 26 Pichincha -0.200 -78.500 7 9 3 3

1757 2 22 Cotopaxi -0.900 -78.600 7 9 1,000 3 3

1797 2 4 12:30 Riobamba -1.640 -78.670 8.3 11 40,000 4 4 4

1808 Quito -0.217 -78.500 1

1859 3 22 13:30 Pichincha -0.300 -78.500 1 3

1868 8 15 19:30 El Angel 0.810 -77.720 8.0 8 1 3 3


Ibarra
1868 8 16 6:30 0.310 -78.180 20 7.7 10 70,000 4 4 4
Colombia S. Pablo
1901 1 7 0 Esmeraldas -2.000 -82.000 25 7.8

1904 11 1 11 Ecuador -1.000 -80.500 7.7

1906 1 31 15:35:51 Off Coast Ecuador 1.000 -81.500 25 8.6 9 1,000 3 2 3

1906 2 7 Colombia-Ecuador 1.000 -81.000

1906 9 28 15:24:54 Ecuador -2.000 -79.000 150 7.9 7 1

1913 2 24 2:30 Gonzanama -3.400 -79.600 50 7.7 7 3 3 3

1914 5 31 Monte Pullerina -0.200 -78.200 10 5.8 10

1923 2 24 0:25 Ecuador -0.400 -78.300 6.8 9 2

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-16
Table 4.2.1 (cont’d). Historic events with M > 6 from 1556 to 2016 (NOOA, 2016).

DATE TIME LOCATION PARAMETERS DEATHS DAMAGE COLLAPSE


DATE TIME LOCATION PARAMETERS DEATHS DAMAGE COLLAPSE
Year Mo Dy UTC Name LAT LON D (km) M MMI No. Dscr.1 $M Dscr.3 No. Dscr. 4
Year Mo Dy UTC Name LAT LON D (km) M MMI No. Dscr. 1 $M Dscr.3 No. Dscr. 4
1924 3 3 8:01 Ecuador -1.600 -78.600 6.9 40 1 1
1924 3 3 8:01 Ecuador -1.600 -78.600 6.9 40 1 1
1933 10 2 15:29:21 Ecuador -2.000 -81.000 33 6.9
1933 10 2 15:29:21 Ecuador -2.000 -81.000 33 6.9
1938 8 10 2:02 Alangasi -0.300 -78.400 10 6.3 10
1938 8 10 2:02 Alangasi -0.300 -78.400 10 6.3 10
1942
1942 55 14
14 2:13
2:13 Guayaquil
Guayaquil -0.750 -81.500
-0.750 -81.500 30
30 7.9
7.9 99 200
200 33 2.5
2.5 22 33
1943
1943 11 30 5:33:03
30 5:33:03 Guayaquil
Guayaquil -2.100 -80.500
-2.100 -80.500 100
100 6.9
6.9 66 22
1949
1949 88 44 19:08
19:08 Pelileo
Pelileo -1.400 -78.500
-1.400 -78.500 10
10 6.7
6.7 10
10
1949
1949 88 55 19:08
19:08 Ecuador
Ecuador -1.500 -78.200
-1.500 -78.200 60
60 6.8
6.8 6,000
6,000 44 7.5
7.5 33
1950
1950 88 55 19:08
19:08 Ecuador
Ecuador -1.500 -78.200
-1.500 -78.200 60
60 6.8
6.8 11
11
1956
1956 11 16
16 23:37
23:37 Ecuador
Ecuador -0.500 -80.500
-0.500 -80.500 7.3
7.3 99 22

1958
1958 11 31
31 9:09
9:09 Ecuador-Colombia -1.500
Ecuador-Colombia -1.500 -79.500
-79.500 7.6
7.6 11
11 11

1960
1960 77 30
30 2:04
2:04 Ecuador
Ecuador -1.500 -79.000
-1.500 -79.000 21
21 11
11 11 11

1969
1969 12 17
12 17 7:13
7:13 Sasquisili
Sasquisili -0.800 -78.300
-0.800 -78.300 27
27 4.2
4.2 33

1976
1976 44 99 7:08:47
7:08:47 Esmeraldas
Esmeraldas 0.780 -79.800
0.780 -79.800 99 6.7
6.7 10
10 11 44 22

1979* 12
1979* 12 31
31 2:59
2:59 Tumaco
Tumaco -1.500 -79.500
-1.500 -79.500 8.2
8.2 300-600
300-600 33

1980
1980 88 18 15:07:53
18 15:07:53 W.Guayaquil
W. Guayaquil -1.948 -80.017
-1.948 -80.017 55
55 5.6
5.6 88 11 55 22

1987
1987 33 66 4:10:44
4:10:44 Napo-Quito
Napo-Quito 0.083
0.083 -77.785
-77.785 10
10 7.2
7.2 5,000
5,000 44 1,500
1,500 44

1987
1987 99 22
22 13:43:38
13:43:38 Ambato
Ambato -0.978
-0.978 -78.050
-78.050 10
10 6.2
6.2 22 11 22 33

1989
1989 66 25
25 20:37:32
20:37:32 Esmeraldas
Esmeraldas 1.134
1.134 -79.616
-79.616 15
15 6.1
6.1 77 22

1990
1990 88 11
11 2:59:55
2:59:55 Pamasqui
Pamasqui -0.059
-0.059 -78.449
-78.449 55 4.4
4.4 44 11 33

1995
1995 10
10 33 1:51:24
1:51:24 Quito
Quito -2.750
-2.750 -77.881
-77.881 24
24 7.0
7.0 22 11 11 83
83 22

1996
1996 33 28
28 23:03:50
23:03:50 Cotopaxi
Cotopaxi -1.036
-1.036 -78.737
-78.737 33
33 5.9
5.9 66 27
27 11 77 33

1998
1998 88 44 18:59:20
18:59:20 Bahia
BahiaDe
DeCaraquez
Caraquez -0.593
-0.593 -80.393
-80.393 33
33 7.2
7.2 33 11 22

2005
2005 11 24
24 23:23
23:23 W.
W.Coast
CoastEcuador
Ecuador -1.364
-1.364 -80.785
-80.785 17
17 6.1
6.1 22 22

2005
2005 10
10 24
24 17:35:34
17:35:34 Baeza
Baeza -0.507
-0.507 -77.745
-77.745 35
35 4.8
4.8 66 11

2007
2007 77 13
13 7:20:26
7:20:26 Zaruma
Zaruma -3.987
-3.987 -79.836
-79.836 50
50 4.5
4.5 11

2007
2007 11
11 16
16 3:13
3:13 Guayaquil
Guayaquil -2.312
-2.312 -77.838
-77.838 123
123 6.8
6.8 11

2009 10 9 18:11:40 Tena -0.962 -77.817 35 5.4 1


2009 10 9 18:11:40 Tena -0.962 -77.817 35 5.4 1
2010 8 12 11:54:16 Manta -1.266 -77.306 207 7.1 1
2010 8 12 11:54:16 Manta -1.266 -77.306 207 7.1 1
2011 10 29 13:50:49 Pomasqui -0.130 -78.370 3 4 1
2011 10 29 13:50:49 Pomasqui -0.130 -78.370 3 4 1
2014 8 12 19:57:58 Quito -0.076 -78.302 5 5.1 2 1
2014 8 12 19:57:58 Quito -0.076 -78.302 5 5.1 2 1
2016 4 16 23:58:37 Muisne, Manabi 0.371 -79.970 19 7.8 662 3 6,998 4
2016 4 16 23:58:37 Muisne, Manabi 0.371 -79.970 19 7.8 662 3 6,998 4
2016 5 18 16:46:11 Manabi 0.465 -79.641 31 6.8 1 1 1 1
2016 5 18 16:46:11 Manabi 0.465 -79.641 31 6.8 1 1 1 1

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-17
Table 4.2.2. Description NOAA database index for death, injuries, economical and house losses.
NOAA DESCRIPTION INDEX
ID CATEGORY UNIT
0 1 2 3 4
1 Deaths
Number 0 1-50 51-100 101-1000 > 1000
2 Injuries
3 Econ. Damage 1990 USD 0 <1M 1-5 M 4-24 M > 25 M
4 Homes Lost
Number 0 1-50 51-100 101-1000 > 1000
5 Homes Damaged

1797, February 4, M s8.3 (MMI XI)


The most powerful earthquake in the history of Ecuador (Chunga et al., 2002) took the life of
more than 40,000 people as per the Comite del Ano Geofisico Internacional del Ecuador (1959).
The event had estimated surface magnitude Ms of 8.3, rupture length of about 70 km (Giesecke et
al., 2004) and MMI Intensity MMI of XI in the epicentral area (Askew & Algermissen, 1985). The
shaking lasted about 3 min (Amodio, 2005) and devastated the city of Riobamba that eventually
relocated ~20 km northeast to the small town of Cajabamba (Kunstaetter & Kunstaetter, 2007).

Figure 4.2.6. The 1797 Riobamba earthquake: Isoseismal curves extrapolated from observed local
intensities, coseismic fissures and cracks, and landslides (Egred, 2000; Baize et al., 2015).
GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity
Report Version 1 4-18
Geotechnical effects were extensive in the form of landslides, cracks, and liquefaction, spread
over large distances, about 200 km in NS and 100 km in EW directions (Fig. 4.2.6). The partial
collapse of the Cushca hill completely buried the western part of the city, overtopped a natural dam
and flooded the plain. The co-seismic landslide (Fig. 4.2.6) had an estimated deposition area of
500×500 m2, with a thickness of 10-30 m with a total volume of 1-3×106 m3 (Baize et al., 2015).

1868, August 15 M s8.0 (VIII) and August 16 M s7.7 (MMI X)


The 1868 earthquakes occurred on August 15th near El Ángel, Carchi province, and on
August 16th near Ibarra in Imbabura province (Giesecke et al., 2004). With estimated surface
magnitudes Ms of 8.0 and 7.7, the earthquakes caused severe damage in NE Ecuador and SW
Colombia, completely ruined the towns of El Ángel and Concepcion and caused up to 70,000
casualties. In the 2nd event every building in Ibarra and nearby Otavalo collapsed and 6,000
people died. In the epicentral area of Imbabura, 15-20,000 people lost their lives. Some
Japanese sources mention a small tsunami at the Ryukyu Islands on August 16 th, but this may
had been caused by the Arica, Peru earthquake of August 13 th, 1868.

1906, January 31, M s8.7, Mw 8.8 (MMI X)

The estimated Mw8.8 megathrust earthquake (Fig. 4.2.7) off the coast of Ecuador-Colombia
(epicenter on Fig. 4.2.3) is among the 10 most powerful events ever recorded (Kanamori,
1977). It generated a strong tsunami reaching 5 m at the shore and killed more than 500 people.

Figure 4.2.7. 1906 Ecuador earthquake photo (Enciclopedia Encarta, Library of Congress).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-19
Kelleher (1972) estimated the rupture length to be ~500 km based on macroseismic data,
extending from 0.5oS, where the Carnegie ridge axis intersects the trench, to 4 oN, at the sharp
bend of the northern Colombian trench. It likely involved the failure of 2-3 large asperities
(Chlieh et al., 2014). Abe (1979) estimated the tsunami magnitude Mt to be 8.7, consistent with
the rupture zone estimate. Actual records from a NS-component Wiechert seismograph at
Göttingen, Germany are shown on Figure 4.2.8 (Kanamori & McNally, 1982).

Figure 4.2.8. Actual records of G2 and G3 waves from the 1906 Colombia-Ecuador earthquake by a
NS-component Wiechert seismograph at Göttingen, Germany (Kanamori & McNally, 1982).

According to USGS (2015), the tsunami was observed all along the coast of Central
America and as far north as San Francisco and west to Japan. The wave arrived in Hilo about
12.5 hours after the earthquake with range of water oscillations of 3.6 m and period 30 min
(Kanamori & McNally, 1982). Based on witness accounts, water level disturbances in
Honolulu Bay began to be observed on February 1 st by a tide gauge about 12 hours after the
earthquake. The tsunami apparently began with a flood; then the oscillations intensified, and
the 4th and highest wave was 0.25 m tall. The max heights of tsunamis from the 1906 and 1979
earthquakes as they arrived in tide stations are shown on Fig. 4.2.17.

1942, May 14, M w7.8 (MMI IX)


This interplate megathrust event appears to have most of its moment release within a 50-
km radius near the epicenter of Fig. 4.2.3 (Swenson & Beck, 1996) and with an unclear number
of failed asperities (Chlieh et al., 2014). Mendoza & Dewey (1984) relocated the epicenter,
including the aftershocks, within the rupture zone of the 2016 Mw7.8 earthquake.

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-20
1949, August 5, M L 6.8 (MMI X)
The 1949 intraplate earthquake was centered near Ambato, the capital of Tungurahua
province (Fig. 4.2.9) and followed an intersection of several NW-SE-trending faults in the
Inter-Andean valley created by the subduction of the Carnegie Ridge. At the time it was the
largest earthquake in the Western Hemisphere for more than 5 years. It was highly destructive
causing the loss of life of approximately 6,000 people (Beauval et al. 2010).

Figure 4.2.9. Topography affected by the 1949 Ambato earthquake (Life magazine, Aug 22, 1949).

About 1/3 of Ambato was destroyed with a 75% of the overall building stock unusable,
including the hospital of Fig. 4.2.10. In the Cathedral 70 people, 60 of which were children,
lost their lives while seeking shelter (Fig. 4.2.11). Fissures on the ground were common (Fig.
4.2.12) and many landslides blocked roads and streams. Water mains and communication lines
were disrupted and the small town of Libertad sank in a gap opened by the earthquake.

Figure 4.2.10. Ambato hospital ruined by the 1949 earthquake (photo by G.E. Lewis, USGS library).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-21
Figure 4.2.11. Ambato Cathedral rescuers work to dig out bodies of 70 people, including 60 children,
who sought shelter in the centuries-old church (photo by Life magazine, 1949).

Figure 4.2.12. Fissures on the ground surface following the 1949 Ambato earthquake (Life, 1949).

The event caused damage in the provinces of Tungurahua, Chimborazo and Cotopaxi and
was felt with intensity IV as far as Cuenca, Guayaquil and Quito. The villages of Guano, Patate,
Pillaro and Pelileo (Fig. 4.2.13) were completely destroyed.
GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity
Report Version 1 4-22
Figure 4.2.13. The village of Pelileo flattened by the Ambato earthquake (Life magazine, 1949).

A cold rain prevented post-earthquake fires, but added to the misery of 100,000 homeless
people who huddled in tents and blankets for shelter and protection of the elements. The
mountainous areas destroyed were some of the nation’s richest food-producing areas (Life,
1949).
1958, January 19, M s 7.6 (MMI IX)
This megathrust earthquake was recorded in 1958 with epicenter shown on Fig. 4.2.3 and
rupture length estimated at 50 km, along one single asperity (Chlieh et al., 2014). This event
also produced a tsunami. Kanamori and McNally (1982) used a recording of this event by a
Press-Ewing seismograph (30 to 90 s) located in Pasadena, CA. The same instrument recorded
the subsequent 1979 event with similar mechanism to estimate the seismic moment of the 1958

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-23
event. As shown on Fig. 4.2.14, the long-period Rayleigh waves have ~1:5.6 ratio between the
1958 and 1979 events, which results to a seismic moment of 5.2×1027 dyn-cm or Mw of 7.7.

Figure 4.2.14. Rayleigh waves recorded by a Press-Ewing seismograph at Pasadena during the 1958
and 1979 events, and used to derive the Mw of the 1958 event (Kanamori & McNally, 1982).

1979, December 12, Mw 8.2, Colombia (MMI IX)


The 1979 megathrust earthquake was the largest in NW S. America since 1942. It initiated
a rupture approximately 280(l)×130(w) km2 in Ecuador (Herd et al., 1981) that propagated NE
along the Colombian coast (Chlieh et al., 2014). The epicenter was just offshore near the port
of Tumaco (Fig. 4.2.3). This event was the last of three that ruptured adjacent parts of the same
megathrust ruptured in the 1906 event (White et al., 2003; Maraillou et al., 2006).

Geotechnical effects were evident, with an estimated 200-km stretch at the coast subsiding
by up to 1.6 m with uplift on the continental slope and land movement that locally disrupted
river drainage. The intense ground shaking (Fig. 4.2.15) caused liquefaction in sand fills and
in Holocene beach, lagoonal, and fluvial deposits (Restrepo & Kjerfve, 2006).

Figure 4.2.15. 1979 Ecuador-Colombia earthquake and tsunami (photos by G. Pararas-Carayannis).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-24
A tsunami swept inland with max observed height of ~6 m at San Juan de la Costa, NE of
Tumaco, causing most of the estimated 300-600 deaths. Figure 4.2.16 shows a tsunami record
from the Esmeraldas port, ~95 naut. mi south of the epicenter. The 1st wave hit the coast ~3
min after the earthquake, with a total of 3-4 waves. The 3 rd wave was the highest and coincided
with low tide, greatly reducing the inundation by 1-3 m and the likely death toll (NGC, 2012).

Figure 4.2.16. Record of 1979 tsunami as recorded in Esmeraldas (Pararas-Carayannis, 1980).

The tsunami was also observed in the Pacific, reaching Hawaii in about 12 hours. Data of
max tsunamis heights as they arrived at tide stations in the 1906 and 1979 events, estimated to
have magnitudes Mt of 8.7 and 8.2, respectively are given in Fig. 4.2.17 (Kanamori & McNally,
1982). The 1979 tsunami was also observed in the Pacific, reaching Hawaii in about 12 hours.
The max wave observed (trough to crest) at Hilo and Kahului was ~40 cm. At Nawiliwili and
Johnston Island the wave was reduced to 10 and 8 cm, while a deep water gauge off the Tokyo
coast did not record wave activity. Figure 4.2.17 presents a computed refraction diagram and
travel times for the 1979 tsunami across the Pacific Ocean (Pararas-Carayannis, 1980).

Figure 4.2.17. Tsunami heights from the 1906 and 1979 earthquakes as they arrived in tide stations
(left; Kanamori & McNally, 1982); refraction diagram and travel time for the 1979 tsunami across the
Pacific Ocean (right; S. Poole, in Pararas-Carayannis, 1980).
The damage by this event (Fig. 4.2.15) and the 1983 Popayán earthquake, led to developing
the 1984 Colombian code for earthquake resistant structures (Garcia, 1984; Garcia et al., 1996).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-25
1987, March 6, Mw 6.1, Mw 7.1 (MMI IX)
The two shallow intraplate earthquakes and occurred in the Andes, far from the subduction
zone, after a month of heavy rains with 600 mm of precipitation (Storchak et al., 2013). The
epicenter was in Napo Province, near the active Reventador volcano (Schuster et al., 1991).
The earthquakes themselves did not cause deaths. Rather, it was the response of the soft
surficial mostly volcanic soils to the strong motions, combined with the heavy rainfall that
generated extensive landslides (Fig. 4.2.18), avalanches and flooding which caused 1,000
deaths and more than 4,000 people missing (ISC-GEM, 2016). Over 800 buildings and 3,000
dwellings collapsed, and ~1,000 buildings and 12,000 dwellings were damaged. These effects
suggest MMI intensity even higher than IX, but this evaluation was complicated due to varying
soil and water table conditions, inconsistent construction practices and scattered population
distributions in the mountains (Schuster & Egred, 1991).

Figure 4.2.18. Looking downstream at the confluence of the Río Malo and the Río Coca, NE Ecuador.
Both channels have been choked by debris flows triggered by the 1987 earthquakes (USGS, 2001).

The major damage due to landslides (Fig. 4.2.19) of the Trans-Ecuadorian (crude oil) and
Poliducto (propane) pipelines and to the highway linking Quito and Lago Agrio (Fig. 4.2.20)
became the largest pipeline damage in history. Fifteen mi were completely destroyed, and 12
mi were seriously damaged (NY Times, 1987), causing an estimated economic loss of 1
billion USD, over 7% of the country’s GDP (Schuster et al., 1991; World Bank, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-26
Figure 4.2.19. Valley wall of Río Malo, NE Ecuador, after the 1987 earthquakes with denudation of
slopes due to avalanches/flows and of valley bottom due to debris flows and flooding (USGS, 2001).

Figure 4.2.20. Route and station locations of the Trans-Ecuadorian pipeline (top) and area of major
damage (Crespo et al., 1987).

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-27
The 1972 Trans-Ecuadorian oil pipeline is composed primarily of 660-mm dia. line pipe
(X-60 grade steel, wall thickness 9.5-20.6 mm) and associated pump and pressure-reducing
stations (Crespo et al., 1991). The pipeline is the main crude oil transportation facility in the
country, conveying virtually all oil from the eastern oil fields to a marine terminal port near
Esmeraldas on the Pacific Ocean and moves 250,000-300,000 barrels of oil daily in a length
of about 500 km. The Poliducto pipeline is a multiproduct facility, including propane gas,
composed of 150-mm dia. line pipe with wall thickness 7.1 mm. It was built after the Trans-
Ecuadorian pipeline and extends from Lago Agrio in the eastern oil field to Quito (Fig. 4.2.20).

Figure 4.2.21. Trans-Ecuadorian pipeline: damage at east bank of the Aguarico River (left); view from
west bank with pipe pulled south by flood with remnant bridge pier and abutment (Crespo et al., 1991).

Figure 4.2.21 shows a segment of the buried pipeline, which was severed at its junction
between the bridge and concrete abutment. Also shown is an east side river view where the
pipeline entered the bridge abutment as an aboveground structure. West of that, the pipeline
was pulled and displaced off its supports for 4 km (Crespo et al., 1987).

1998, August 4, Mw7.1 (VIII)


The 1998 Bahía de Caráquez megathrust earthquake was located 10 km north of Bahía de
Caráquez at a sandy peninsula in the western coast of Ecuador and likely ruptured the asperity
directly south of the Equator (Chlieh et al., 2014). In 1998 the whole coast of Ecuador was
affected by the El Niño phenomenon.

Bahía was the most severely affected by the torrential rains and mudslides, leaving the
town isolated and without basic services and food supplies. Shortly after El Niño, the town was
struck by the Mw7.1 earthquake. Many buildings fell to the ground (Fig. 4.2.22) and almost all
were severely damaged (Castro, 2007). It was felt strongly throughout Ecuador, especially in

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-28
Guayaquil and Quito, and in Cali, Colombia. In Bahía, electricity, telephone, and water
services disrupted, and many buildings were damaged (Segovia et al., 1999).

Figure 4.2.22. Building Calypso, before and after the Bahía de Caráquez earthquake (Castro, 2007).

This event better defined the Bahía Earthquake Zone (BEZ in pink on Fig. 4.2.3) by
studying its aftershocks (Font et al., 2013). The BEZ shows a characteristic behavior of an M~7
event every 50-100 years and seems to behave independently of the northern megathrust as
indicated by the additional 1956 Mw6.95 and 1896 M w7 with epicenters south of the 1906 and
1942 rupture zones.

2016, April 16, Mw7.8(VIII)


Details on the M w7.8 earthquake near Muisne in the Manabi Province at the west coast of
Ecuador will be presented in the following chapters, including recorded ground motions of the
main event.

GEER-ATC Muisne, Ecuador, Earthquake 4-Seismotectonics and Historic Seismicity


Report Version 1 4-29
Blank Page
CHAPTER 5
Strong Ground Motions

PGA = 1.4 g

GEER-ATC Muisne, Ecuador 2016


Report Version 1

 
Blank Page
5.1 Seismological Aspects
The seismological aspects of the main M w7.8 event of the April 16th 2016 were provided
by the GEER-ATC collaborating agency, the Instituto Geofísico at the Escuela Politécnica
Nacional (IG-EPN), translated as the Ecuadorian Geophysical Institute from the National
Polytechnic School (Singaucho et al., 2016). This subduction event is the largest interface
thrust earthquake since the 1979 Mw8.1 megathrust event that activated the northernmost
segment of the 1906 rupture zone (see Chapter 4).

The seismological parameters as interpreted by IG-EPN (2016), the United States


Geological Survey, USGS (2016), and the Global Centroid-Moment-Tensor (CMT) catalog
(Dziewonski et al., 1981, Ekström, G. et al., 2012) are shown on Table 5.1.1. The focal
mechanism according to the Global CMT catalog is presented on Figure 5.1.1.

Table 5.1.1. Seismological parameters for the main 4/16/16 event by IG-EPN, USGS, and CMT.

CATALOG DATE UNIV DEPTH MAGN


LAT LON
(agency) (mm/dd/yy) TIME (km) Mw
IG-EPN 0° 21' 00" N 80°09'36"W 17 7.4
USGS 4/16/2016 23:58 0° 22' 55" N 79°55'19"W 21 7.8
CMT 0° 09' 36" S 80°20'60"W 24 7.8

Figure 5.1.1. Focal mechanism for the main 4/16/216 event reported in the Global CMT catalog
(Dziewonski et al., 1981; Ekström, G. et al., 2012).

The epicenter of the main Mw7.8 event and the geographical distribution of the epicenters
of its major aftershocks are shown on Figure 5.1.2 along with epicenters of historic past events.
The seismological characteristics of the major aftershocks are presented on Table 5.1.2 (IG-
EPN, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-1
Figure 5.1.2. Epicenters of the main event (purple star) and its aftershocks (circles colored and sized
according to magnitudes). Epicenters of historical earthquakes are shown in yellow stars and
background seismicity prior to April 16 th is shown as black and white circles.

The aftershocks bounded an approximate subduction segment of 100 km long by 80 km


wide which corresponds to the rupture area along the interface. The resulting 8,000 km2 of
interface rupture coincides with the outcome obtained from the source scaling relations for
interface subduction-zone earthquakes developed by Strasser et al. (2010).

The WNW-ESE aftershocks alignment immediately north of the 2016 epicenter (Fig. 5.1.2)
appear to be the expression of an efficient barrier for the northern propagation of the rupture.
It is appealing that there is almost no post-earthquake seismicity north of this lineament. The
nature of the northern barrier is not known. Several attempts to identify controlling structures

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-2
with the aid of geophysical and bathymetric cruises did not find strong evidence (Collot et al.,
2004) for a segment boundary, although a pervasive seismicity with the same location and
orientation has been present since the beginning of the instrumental seismic monitoring by the
IG-EPN starting in the early 90’s (Fig. 5.1.3, Yepes et al., 2016).

Table 5.1.2. Seismological parameters for major aftershocks of the 4/16/16 event (IG-EPN, 2016).

DATE UNIV DEPTH MAGN


LAT LON
(mm/dd/yy) TIME (km) Mw
04/17/16 21:35 0° 54' 36" S 80° 33' 36" W 10 6.5
04/20/16 8:35 0° 40' 48" N 80° 13' 12" W 4 6.3
04/22/16 3:03 0° 10' 48" S 80° 46' 12" W 10 6.2
05/18/16 7:57 0° 26' 24" N 79° 57' 0" W 7 6.6
05/18/16 16:46 0° 27' 36" N 79° 50' 24" W 9 6.7
07/11/16 2:01 0° 35' 24" N 79° 46' 12" W 10 6.2

The southern rupture boundary is also marked by a less conspicuous WNW-ESE alignment
of aftershocks immediately south of the 1942 epicenter shown on Fig. 5.1.2). This
concentration lies immediately north of the loci of background seismicity and aftershocks of
the Mw7.1 1998 BEZ earthquake as shown on Fig. 5.1.2 (Yepes et al., 2016). The apparent
seismic gap between the two alignments (Fig. 5.1.3) was filled with the 2016 earthquake.

Figure 5.1.3. The 110 years of instrumental seismicity in Ecuador, 1900–2009. Earthquakes with M w
≥7 are plotted as stars (Yepes et al., 2016).

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-3
Seismic activity that appears on Fig. 5.1.2 south of the BEZ is not be part of the aftershocks,
since it was present before the April 16th main event (see Fig. 5.1.3). It is rather related to the
slow slip activity found below La Plata Island (Vallée et al., 2013).

More refined studies modeling near field geodetic, seismological and accelerometric data
(Nocquet et al., 2016) indicate that the 2016 earthquake ruptured two asperities along an 80-
100 km by 40-50 km area of the interface. In both models (the geophysical and geodetic data
inversion and the aftershock distribution), the area where the seismic energy was released
during the main event of 4/16/16 seems to overlap with the area ruptured in the 1942 event. Of
great interest is the fact that a southward directivity of the released seismic energy observed
(Nocquet et al., 2016) is consistent with the large damage observed south of the rupture area.
The same modeling did not find significant co-seismic slip, which may explain the lack of a
significant tsunami following the earthquake.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-4
5.2 Strong Motion Records
Strong ground motion acceleration time histories records were provided by the Instituto
Geofísico at the Escuela Politécnica Nacional (IG-EPN), translated as the Ecuadorian
Geophysical Institute from the National Polytechnic School (Singaucho et al., 2016). IG-EPN
manages and maintains the National Accelerometer Network RENAC (Red Nacional de
Acelerógrafos) which has digital accelerometers throughout Ecuador, as shown on Fig. 5.2.1.
The data used in this work were obtained from RENAC of IG-EPN. The installation and
expansion of RENAC as operated by the IG-EPN was made possible utilizing funds
from the Escuela Politecnica Nacional, part of Proyecto SENESCYT PIN -08-
EPNGEO-0001 “Fortalecimiento del Instituto Geofísico: Ampliación y Modernización del
Servicio Nacional de Sismología y Vulcanología,” and of investment project termed
“Generación de Capacidades para la Difusión de Alertas Tempranas y para el Desarrollo
de Instrumentos de Decisión ante las Amenazas Sísmicas y Volcánicas dirigidos al Sistema
Nacional de Gestión de Riesgos.” During earthquakes, sensors can be susceptible to power
outages or physical damage to components (e.g., solar panels) that affect their proper operation.

Figure 5.2.1. Spatial distribution of the RENAC and OCP accelerometer stations. The epicenter of the
main event is shown as a star, and the circles at selected stations reflects the Peak Ground Acceleration
(PGA) recorded in m/s2 (by IG-EPN, Singaucho et al., 2016).

Additional instruments are available from OCP (oil pipeline network) and LMI (a
collaborative project between IG-EPN and the Institute of Research for Development, IRD, of
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-5
France). Details on all instruments are given presented on Table 5.2.1, including the recorded
Peak Ground Acceleration (PGA) of the three components (E, N, Vertical) and the epicentral
distance from the main April 16th, 2016 event (Singaucho et al., 2016).

Table 5.2.1. Accelerometer stations in Ecuador owned by RENAC (maintained by IG-EPN, Singaucho
et al., 2016), OCP, and LMI with PGA values recorded during the main event.
Geographic Epicentral
Elevation Instrument VS30 PGA (g)
Station City Coordinates Distance
Type
Latitude Longitude (m) (km) (m/s) EW NS VER
AAM2 Ambato -1.269 -78.611 2,664 Guralp 235 - 0.026 0.035 0.015

ACH1 Machala -3.287 -79.910 13 Reftek 407 - 0.025 0.024 0.008

ACHN Chone -0.698 -80.084 18 Reftek 120 200 0.330 0.370 0.173

ACUE Cuenca -2.910 -78.959 2,578 Reftek 381 - 0.036 0.030 0.018

PDNS Pedernales 0.111 -79.991 442 Kephren 29 - 1.061 0.974 0.573

APED Pedernales 0.068 -80.057 15 Reftek 36 342 1.408 0.829 0.742

AV18 Quininde 0.313 -79.478 107 Reftek 52 - 0.144 0.132 0.069

AV21 Viche 0.661 -79.547 62 Reftek 54 - 0.193 0.150 0.091

AMA1 Esmeraldas 0.935 -79.725 234 Reftek 67 - 0.426 0.203 0.135

AES2 Esmeraldas 0.991 -79.646 4 Reftek 76 - 0.154 0.110 0.044

APR2 Puerto Quito 0.077 -78.968 804 Reftek 113 - 0.101 0.110 0.057

ASDO Santo Domingo -0.263 -79.124 615 Guralp 115 - 0.206 0.111 0.051

ALOR San Lorenzo 1.293 -78.847 22 Reftek 159 - 0.026 0.027 0.015

APO1 Portoviejo -1.038 -80.460 47 Reftek 167 224 0.318 0.380 0.104

AMNT Manta -0.941 -80.735 38 Guralp 171 496 0.405 0.524 0.165

EPNL Quito -0.212 -78.492 2,813 Guralp 174 - 0.027 0.020 0.013

AV11 Guallabamba -0.073 -78.371 2,058 Reftek 181 - 0.030 0.039 0.020

AOTA Otavalo 0.240 -78.256 2,529 Guralp 188 - 0.043 0.035 0.019

AIB1 Ibarra 0.347 -78.125 2,208 Guralp 202 - 0.049 0.058 0.012

AIB2 Ibarra 0.349 -78.106 2,298 Guralp 204 - 0.021 0.033 0.009

ALAT Latacunga -0.926 -78.618 2,777 Guralp 206 - 0.032 0.028 0.012

APS4 Papallacta -0.371 -78.106 2,887 Reftek 220 - 0.007 0.006 0.002

ATUL Tulcán 0.772 -77.723 3,097 Guralp 251 - 0.016 0.021 0.007

AGYE Guayaquil -2.054 -79.952 30 Guralp 270 1800 0.019 0.023 0.015

AMIL Milagro -2.181 -79.529 20 Guralp 288 - 0.052 0.046 0.019

ALIB La Libertad -2.243 -80.846 40 Guralp 308 429 0.042 0.040 0.021

ALJ1 Loja -3.987 -79.197 2,147 Reftek 492 - 0.015 0.016 0.009

GYE1 Guayaquil -2.251 -79.910 7 Reftek 292 178 0.059 0.070 0.020

GYE2 Guayaquil -2.199 -79.899 11 Reftek 286 101 0.094 0.097 0.038

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-6
The two types of instruments, Guralp and Reftek in Table 5.2.1, are shown on Fig. 5.2.2.
The 51 installed Guralp CMG-5TDE instruments have an internal memory storage of 8 GB
and the 28 installed Reftek 130-SMA instruments have two 2 discs of 8 GB storage each. Both
instruments can transmit data in real time (EG-IPN, 2016).

Figure 5.2.2. The two main types of strong motion instruments Guralp CMG-5TDE (left) and Reftek
130-SMA (right). Details can be found in EG-IPN (2016s).

A total of 30 three-component uncorrected digital acceleration time series were provided


from the main event. Most of records were from stations of RENAC, with some from OCP (oil
pipeline network) or LMI (a collaborative project between IG-EPN and the Institute of
Research for Development, IRD, France). Some RENAC stations were adversely affected, as
power outages and building collapses relegated many of the sensors as useless. The IG -EPN
Informe Especial No. 18 stations worked well and their records were disseminated.

All provided records were processed by the GEER team following the PEER standard
procedure by Ancheta et al. (2013), which includes inspection of record quality, selection of
time windows, such as P-waves, S-waves, and coda waves, and component specific filter
corner frequencies to optimize the usable frequency range. From the 30 stations, only one
station (PRAM) was rejected due to the quality of signal and is not listed on Table 5.2.1.

Table 5.2.2 presents details of 10 selected stations that recorded the main event with
location (nearest city, latitude, longitude), closest distance to the rupture plane (R RUP) using the
USGS (2016) finite fault solution, recorded PGAs, and Vs30 if available in Vera-Grunauer
(2014), or directly measured by Geoestudios (2016) during the GEER-ATC mission.

Figure 5.2.3 shows the USGS Intensity ShakeMap together with the projected rupture plane
to the surface and Strong Motion Stations (SMS) locations, color-coded according to the geo-
mean of the two horizontal components of PGA. In general, both the ShakeMap and SMS show
that the shaking was more intense towards the south of the rupture plane. Also, both types of
intensity tools (ShakeMap and SMS) indicate site effects in the alluvial soft deposits of
Guayaquil, with closest distance to the rupture plane (R RUP) between 160 and 180 km.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-7
Table 5.2.2. Summary of 10 of the stations with ground motion records processed by GEER.

Geographic Coordinates RRUP VS30 PGA (g)


Station City
(km)
Latitude Longitude (m/s) EW NS VER
PDNS Pedernales 0° 6' 39.6" N 79° 59' 27.6" W 21 - 1.034 0.942 0.573

APED Pedernales 0° 4' 4.8" N 80° 3' 25.2" W 20 342a 1.408 0.83 0.742

AES2 Esmeraldas 0° 59' 27.6" N 79° 38' 45.6" W 51 - 0.154 0.111 0.044

ACHN Chone 0° 41' 52.8" S 80° 5' 2.4" W 34 200a 0.328 0.371 0.173

APO1 Portoviejo 1° 2' 16.8" S 80° 27' 36" W 73 224a 0.317 0.381 0.105

AMNT Manta 0° 56' 27.6" S 80° 44' 6" W 76 496a 0.404 0.525 0.162

EPNL Quito 0° 12' 43.2" S 78° 29' 31.2" W 104 - 0.027 0.02 0.013

AGYE Guayaquil 2° 3' 14.4" S 79° 57' 7.2" W 155 1800b 0.019 0.024 0.015

AGY1 Guayaquil 2° 15' 3.6" S 79° 54' 36" W 175 178b 0.059 0.065 0.02

AGY2 Guayaquil -2° 11' 56. 4" S 79° 53' 56.4" W 170 101b 0.094 0.098 0.038
a Measured by Geoestudios (2016). b Measured by Vera-Grunauer (2014).

The strongest motion with PGA = 1.41 g was recorded in the EW component of the APED
Pedernales station, with R RUP of around 20 km and Vs,30 of 342 m/s (Geoestudios, 2016) and
PGA = 0.83 g in the NS component. Another Pedernales station, PDNS (R RUP = 21 km),
recorded PGAs of 1.03 and 0.94 g in the EW and NS components. Figure 5.2.4 shows EW
acceleration time series for selected stations, overlying a map of Ecuador. Station AV21, near
Esmeraldas has a smaller distance than station ACHN (25 vs. 34 km), but recorded half the
amplitude of ACHN. Stations APED, PDNS, AMNT, ACHN and APO1 recorded PGA > 0.3
g. AGYE and AGY2 recorded significantly different PGAs (0.02 vs. 0.094g) as they are based
on hard rock (VS30 = 1,800 m/s) and soft soil (VS30 = 101 m/s), respectively.

Even though AMNT has longer and similar RRUP to ACHN and APO1, its PGA in EW
direction is much larger (0.5 vs. 0.33 and 0.3 g, respectively). During reconnaissance, we
observed that AMNT was near a cliff, which may imply topographic effects. By visual
inspection of the records, higher frequency contents are dominant for most records. Exceptions
are stations ACHN and AGYE2, with lower frequency content, probably due to site effects.

Long-period motions are better observed in terms of velocity or displacement time


histories. A comparison of acceleration, velocity and displacement time histories for stations
APED, AMNT, ACHN, and APO1 are plotted on Fig. 5.2.5 in the same scale.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-8
Figure 5.2.3. Contours of Intensity (ShakeMap) from USGS and Strong Motion Stations (SMS), color-
coded by the intensity of the motion in terms of the geo-mean Peak Ground Acceleration (PGA).

Figure 5.2.4. Acceleration time series in the EW component overlying a map of Ecuador with main
cities and recording stations. Colored circles same as in Fig. 5.2.3.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-9
Figure 5.2.5. Acceleration, velocity and displacement time histories for the EW component of stations
APED, AMNT, ACHN and APO1.

High-frequency content is evident at the APED from visual inspection of the velocity time
series. AMNT and APO1 show lower frequency content from 2 to 3 s. At ACHN, where the
acceleration time series showed lower frequency content (higher periods) than AMNT and
APO1, but higher frequency than the same two stations in velocities, say between 1 and 2 s.

Acceleration, velocity and displacement time histories for the 29 stations of Table 5.2.1
with their 3 components are plotted in Figs 5.2.6 to 5.2.34 showing Husid plots, 5%-damped
pseudoacceleration response spectra (Sa) and ground motion parameters including PGA, Peak
Ground Velocity (PGV), Peak Ground Displacement (PGD), significant durations (D5-75 and
D5-95), Arias intensity (Ia) and mean period (T m), with information about seismic stations and
processing of the record. The processed time histories are available for download from the
GEER website www.geerassociation.org.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-10
Figure 5.2.6. Recorded motions of the main Mw7.8 April 16th, 2016 event: Station AAM2, Ambato.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-11
Figure 5.2.7. Recorded motions of the main M w7.8 April 16th, 2016 event: Station ACH1, Machala.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-12
Figure 5.2.8. Recorded motions of the main M w7.8 April 16th, 2016 event: Station ACHN, Chone

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-13
Figure 5.2.9. Recorded motions of the main M w7.8 April 16th, 2016 event: Station ACUE, Cuenca.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-14
Figure 5.2.10. Recorded motions of the main M w7.8 April 16th, 2016 event: Station PDNS, Pedernales.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-15
Figure 5.2.11. Recorded motions of the main Mw7.8 April 16th, 2016 event: Station APED, Pedernales.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-16
Figure 5.2.12. Recorded motions of the main M w7.8 April 16th, 2016 event: Station AV18, Quininde.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-17
Figure 5.2.13. Recorded motions of the main M w7.8 April 16th, 2016 event: Station AV21, Viche.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-18
Figure 5.2.14. Recorded motions of main M w7.8 April 16th, 2016 event: Station AMA1, Esmeraldas.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-19
Figure 5.2.15. Recorded motions of the main M w7.8 April 16th, 2016 event: Station AES2, Esmeraldas.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-20
Figure 5.2.16. Recorded motions of Mw7.8 April 16th, 2016 event: Station APR2, Puerto Quito.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-21
Figure 5.2.17. Recorded motions of Mw7.8 April 16th, 2016 event: Station ASDO, Santo Domingo.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-22
Figure 5.2.18. Recorded motions of Mw7.8 April 16th, 2016 event: Station ALOR, San Lorenzo.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-23
Figure 5.2.19. Recorded motions of the M w7.8 April 16th, 2016 event: Station APO1, Portoviejo.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-24
Figure 5.2.20. Recorded motions of the M w7.8 April 16th, 2016 event: Station AMNT, Manta.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-25
Figure 5.2.21. Recorded motions of the M w7.8 April 16th, 2016 event: Station EPNL, Quito.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-26
Figure 5.2.22. Recorded motions of the M w7.8 April 16th, 2016 event: Station AV11, Guallabamba.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-27
Figure 5.2.23. Recorded motions of the M w7.8 April 16th, 2016 event: Station AOTA, Otavalo.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-28
Figure 5.2.24. Recorded motions of the M w7.8 April 16th, 2016 event: Station AIB1, Ibarra.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-29
Figure 5.2.25. Recorded motions of the M w7.8 April 16th, 2016 event: Station AIB2, Ibarra.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-30
Figure 5.2.26. Recorded motions of the M w7.8 April 16th, 2016 event: Station ALAT, Latacunga.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-31
Figure 5.2.27. Recorded motions of the M w7.8 April 16th, 2016 event: Station APS4, Papallacta.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-32
Figure 5.2.28. Recorded motions of the M w7.8 April 16th, 2016 event: Station ATUL, Tulcán.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-33
Figure 5.2.29. Recorded motions of the M w7.8 April 16th, 2016 event: Station AGYE, Guayaquil.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-34
Figure 5.2.30. Recorded motions of the M w7.8 April 16th, 2016 event: Station AMIL, Milagro.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-35
Figure 5.2.31. Recorded motions of the Mw7.8 April 16th, 2016 event: Station ALIB, La Libertad.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-36
Figure 5.2.32. Recorded motions of the M w7.8 April 16th, 2016 event: Station ALG1, Loja.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-37
Figure 5.2.33. Recorded motions of the M w7.8 April 16th, 2016 event: Station GYE1, Guayaquil.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-38
Figure 5.2.34. Recorded motions of the M w7.8 April 16th, 2016 event: Station GYE2, Guayaquil.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-39
COMPARISON TO GROUND MOTION PREDICTION EQUATIONS (GMPEs)

The SMS ground motions were compared to the subduction Ground Motion Prediction
Equation (GMPE) BCHydro (Abrahamson et al., 2016) with parameters presented in Table
5.2.3. The closest distance to the rupture RRUP was estimated using the USGS finite fault model.
An average shear wave velocity in the upper 30 m (Vs,30) of 400 m/s was selected for those
stations where measurements were not available. Although few stations are located in the back -
arc (towards the east of Andean mountains), where more attenuation occurs, in this evaluation
we assumed that all stations are in the fore-arc region. The assumptions of VS30 = 400 m/s and
all stations being in the fore-arc region should not significantly affect the results. However, a
different finite fault model may change the estimation of R RUP and which may change the
distance and thus the GMPE predicted values.

Table 5.2.3. Parameters for the BCHydro GMPE (Abrahamson et al., 2016). Stations were assumed to
be in the fore-arc, although some far stations are in the back-arc, east of Andeans mountains.

PARAMETER VALUE COMMENTS


Magnitude, MW 7.8 --
Ftype 0 0 (interface), 1 (intra-slab)
1
Faba 0 0 (force-arc), 1 (back-arc )
Flag_deltaC1 1 Period-dependent values of ΔC1
Period (T) several --
Distance to Rupture (RRUP) see Table 5.2.1 USGS finite fault model
Hypocentral Distance (Rhyp) -- intra-slab events only
VS,30 see Table 5.2.1 400 m/s if Vs,30 not known

The recorded PGA and Sa at periods T of 0.2, 1.0 and 3.0 s were compared to the median
and plus/minus one standard deviation (±) of the GMPE predictions. The between-event
residual (δB) for these periods were also calculated. Figure 5.2.35 shows GMPEs and recorded
geometric-mean Sa for each station along with the median curve for the main event adjusted
by δB. Calculated between-event residuals were -0.54, -0.61, -0.105 and -0.035 (in natural log
units) for PGA, Sa (0.2 s), Sa (1.0 s) and Sa (3.0 s), respectively. For short spectral periods T,
the median recorded Sa is closer to the median  of the BCHydro GMPE. For longer T, the
median recorded Sa is similar to the median from other subduction earthquakes. Figure 5.2.36
shows PGA δB for all earthquakes used in the BCHydro GMPE regression together with the
2010 Maule Chile and 2011 Tohoku, Japan earthquakes. The δB for the 2016 Muisne, Ecuador
earthquake is within expected values of between-event residuals.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-40
For the same periods (PGA, 0.2, 1.0, 3.0 s), the within-event residuals (δWes) were
estimated for each station and plotted against R RUP on Fig. 5.2.37. The standard deviation (φ)
of the δWes for each period is between 0.8 and 1.12, a range of values larger than the GMPE
φ of 0.6, which may be explained by the following reasons:

 Uncertainty in R RUP distance: A different finite fault model will likely change R RUP and
move the data points closer or away from each other.
 Distance attenuation term: The distance attenuation term is different between regions and
the GMPE uses a global attenuation term.
 Bias of stronger ground motions: At larger distances, some instruments may have not been
triggered, resulting in a bias of stronger motions in stations with large R RUP.
 Site effects: For most stations, site effects were not incorporated for lack of data. For long
periods, site effects may be important and thus increase the  of the within-event residuals.

Figure 5.2.35. Observed ground motions (green dots), predicted by the BCHydro GMPE median (red
solid line), median ± (red dashed line), and median curve for main Ecuador event (blue dashed line).

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-41
Figure 5.2.36. Between-event residuals for PGA showing points used for the BCHydro GMPE
regression with Maule, Chile (2010), Tohoku, Japan (2011), and Muisne, Ecuador (2016) earthquakes.

Figure 5.2.37. Within-event residuals (δWes) for PGA and Sa at periods of 0.2, 1.0 and 3.0 s.

A comparison of the recorded 5%-damped geometric-mean Sa with predicted median and


median ± values from the BCHydro GMPE are presented on Fig. 5.2.38 which highlights
ground motion characteristics from selected stations. In Pedernales, the highest intensity area,
Sa values from strong motion records in stations PDNS and APED are in agreement with
median values for periods T longer than 1 s, but closer to the median  for T shorter than 1 s.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-42
Figure 5.2.38. Comparison of GMPE predicted 5% damped median and +/- 1 standard deviation
response spectra (red lines) with recorded geo-mean response spectra (blue lines) for stations (a) APED,
(b) PDNS, (c) AMNT, (d) APO1, (e) ACHN, (f) AV21, (g) AGYE, and (h) AGY2.

AMNT records show higher Sa than the median + throughout the entire range of periods.
After analyzing the geometry of the slopes near the stations and in-situ Vs measurements, the
elevated Sa in the short-period range (< 0.5 s) may be explained by site effects due to the
presence of a thin, low-Vs surficial layer and deeper rock layer identified throughout the site
with possible contribution from topographic effects. However, although AMNT is about 5 m
away from a 24-m tall nearly vertical cliff and has a northerly aspect, topographic effects may
not be the main cause of the higher than expected recorded motions because the waves
propagated from north to south, traveling away rather than into the slope face (Ashford & Sitar,
1997). Also, the orientation of the long axis of the slope (azimuth of 80 o) would suggest
topographic effects in the NS component, which is not supported by the recorded motions.

At period T ~ 3 s, the AMNT, APO1 and ACHN records show a slight bump in spectral
acceleration with Sa values greater than the median +σ. These Sa peaks at the three stations
suggest possible directivity or path effects, which can be further investigated using aftershocks

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-43
records. In the short period range, APO1 shows a significant Sa peak at 0.5 s, with amplitude
5 times larger than the PGA and almost double the median +. ACHN also has a clear Sa peak
at 1.5 s with amplitude approximately 5 times larger than the PGA and well above the median
+. Station AV21 has smaller R RUP than ACHN, but shows less intensity with longer duration
(close to median). This trend is consistently observed between stations at the north (AMA1
and AV21) and stations at the south of the rupture plane (AMNT, ACHN and APO1).

At Guayaquil, stations AGYE and AGY2, on rock and soft soil respectively, show clear
differences in spectral shapes. Sa values at AGYE are similar to the median  of the GMPE,
while Sa at AGY2 is between the median and median +. The spectral shape at AGY2 does
not show a clear peak, but rather has an entire zone of high Sa values ~0.2 to 0.3 g for T
between 0.3 and 2.0 s, which is likely indication of soft soil site effects (Vera-Grunauer, 2014).

COMPARISON TO CODE-BASED DESIGN SPECTRA (NEC-15)

The Ecuadorian Construction Norm (NEC-15) has been in effect since 2015. The code
includes Chapter 2, “Seismic Hazard and Seismic Resistant Design,” with a design spectrum
shaped as shown on Fig. 5.2.39 and Sa values calculated based on the location, intensity of
shaking, and site classification. The design seismic event has a return period Tr of 475 years
(or a probability of exceedance of 10% in 50 years).

Figure 5.2.39. Map of NEC-15 “Z” factor, equal to the PGA in rock or stiff soil conditions, for a return
period of 475 years or 10% probability of exceedance in 50 years (NEC, 2015).

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-44
The NEC-15 design spectrum is constructed based on the “Z” factor, which corresponds to
the PGA for rock or stiff soil conditions for the 475-year event. The value of “Z” is mapped on
Fig. 5.2.39. Almost the entire coastline is mapped with Z = 0.5 g, which is a deterministic cap,
since probabilistic PGA values where higher. The code officials decided to limit Z to 0.5 g due
to construction cost implications. Once Z is obtained, site amplification factors (F a, Fd , Fs) are
derived by classifying the site as per ASCE 7-10 procedures using geotechnical data for the
top 30 m of overburden. Requirements for site-specific studies are given for special cases, such
as liquefiable sites or sites with thick soft and plastic clay deposits. Other structural factors,
such as η and r, are incorporated to produce the design acceleration spectrum of Fig. 5.2.40.

Figure 5.2.40. Shape of design elastic acceleration spectra from NEC-15.

The NEC-15 design spectra derived for the sites of stations APED, PDNS, AMNT, APO1,
ACHN, and AGY2 were compared to actual geomean response spectra from the main event.
The parameters used to derive the code spectra are provided on Table 5.2.4. Site class was
assigned for stations where in-situ VS,30 values were available at the time of writing this report.

Table 5.2.4. Parameters for constructing the NEC-15 design acceleration response spectra for stations
APED, PDNS, AMNT, APO1, ACHN, and AGY2 .

Station Z Site Class VS30 ƞ r Fa Fd Fs


APED 0.5 D 342 1.8 1.0 1.12 1.11 1.40

PDNS 0.5 C 400 1.8 1.0 1.18 1.06 1.23

AMNT 0.5 C 496 1.8 1.0 1.18 1.06 1.23

APO1 0.5 D 224 1.8 1.0 1.18 1.11 1.40

ACHN 0.5 D 200 1.8 1.0 1.12 1.11 1.40


1
AGY2 0.4 F 101 1.8 1.5 1.00 1.60 1.90
1
Although this site is classified as “F”, it will be compared site “E” NEC-15 design spectra for Guayaquil.
GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions
Report Version 1 5-45
An exception is the Guayaquil AGY2 station, for which the zonation study by Geoestudios
(2014) indicates presence of 3 m of highly organic clay and very thick (> 30 m) soft to medium
stiff clays, which makes this site class “F.” For comparison purposes, class “E” design values
will be used and compared to site-specific spectra (Argudo et al., 2001; Geoestudios, 2014).

Figure 5.2.41. Comparison of recorded geometric-mean, median GMPE predictions, and NEC-15
design spectra for stations (a) APED, (b) PDNS, (c) AMNT, (d) APO1, (e) ACHN and (f) AGY2.

Figure 5.2.41 compares recorded geometric-mean Sa and median Sa from BCHydro GMPE
with NEC-15 design Sa. The APED and PDNS spectra (Fig 5.2.41.a and b), are well above
design, while AMNT (Fig. 5.2.41.c) has Sa values comparable to NEC-15. However, AMNT
station had greater than median+ values compared to the GMPE at the entire range of periods,
indicating that site or topographic effects may have played an important role in this station. For
the APO1 station (Fig 5.2.41.d) two design spectra are shown. The black line is the actual
design spectra considering the value of “Z” obtained from Figure 5.2.39, which includes a cap
of 0.5 g. The gray design spectrum in Fig. 5.2.41f considers the “Z” value without the 0.5 g
limit, obtained from hazard curves in the NEC-15 for the capitals of the provinces. The results
are similar, with the design spectra being higher for almost every period with the exception of
the high peak of 1.6 g at ~0.5 s, where the recorded Sa are higher than design levels by a factor

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-46
of 1.3 and 1.6 for the case with and without the limit. Structures near this station with T ~ 0.5
s may have suffered significant damage. ACHN (Fig. 5.2.41e) shows recorded Sa lower than
design for T < 1 s, and exceedance of design between 12.5 s with a clear peak at ~1.5 s and
Sa values twice the design ones in this period range. The AGY2 (Fig. 5.2.41f) Sa are below
design levels (assuming class “E”) and site-specific studies in Guayaquil. Even though the Sa
amplitude differs, the spectral shapes are similar to the code and site-specific studies.

SUMMARY AND FILES OF STRONG MOTION RECORDINGS

Acceleration time histories of 87 recordings from 29 stations with 3 components from the
main Mw7.8 April 16th, 2016 event have been processed for this report following the PEER
standard data processing method by Ancheta et al. (2013). The data were compared to values
derived using subduction earthquake GMPE and the Ecuadorian Construction Norm (NEC-
15). The median values of recorded Sa for this event in general fall below the medians values
from other subduction events in short periods. At longer periods, the median Sa of this event
is similar to the median from other subduction earthquakes. Two recordings show large
intensities with PGA on the order of 1 g in the town of Pedernales. Seismic stations located
towards the south of the rupture plane recorded larger shaking with shorter duration compared
to the stations located towards the north, with similar closest distance to the rupture plane
(RRUP). Topographic effects were likely present at seismic station in AMNT Manta station,
based on the reconnaissance field investigation and frequency contents of the recorded motion.
As a result, recorded Sa values were well above the median values from the GMPE and similar
values to the NEC-15 design spectra at the AMNT site. Stations ACHN and APO1, located in
Chone and Portoviejo, respectively, show clear peaks at spectral periods of 1.8 and 0.5 s,
respectively, with higher Sa values at these periods than the design ones, while at all other
periods, the Sa values are below design levels. Strong site effects were observed at stations
AGYE and AGY2 in Guayaquil city, located on rock and soft soil, respectively, with
amplification factors of ~ 5 for PGA and ~10 for Sa between 0.2 and 2 s. The processed time
histories can be downloaded from the GEER web site www.geerassociation.org/.

AFTERSHOCKS

Selected recordings from the aftershocks have been provided to our team and are being
processed using the PEER procedure. They will be available in the next version of this report.

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-47
5.3 Tsunamis
The main Mw7.8 earthquake of April 16th, 2016 in Ecuador triggered a tsunami, which was
registered by the NOAA 32411 and 32413 DART buoys in the Panama and Peru basins,
respectively. The registrations happened 2 and 3 hours after the earthquake, when the waves
passed over the positions of the buoys. Furthermore, the seismic waves propagated over the
seabed, affecting the register at the pressure sensor of these buoys until one hour after the
earthquake (Fig. 5.3.1). Roughly, the observations indicate that the period of the tsunamis has
been about 40 min with an amplitude close to 1 cm (Toulkeridis et al., 2016).

Figure 5.3.1. Tsunami water levels from the main earthquake registered at NOAA 32411 (top) and
32413 (bottom) DART buoys in the Panama and Peru basins, respectively. Information by NOAA /
PMEL / Center for Tsunami Research (NOAA, 2016).

Locally, the tsunami was registered immediately by the INOCAR-DART buoy which is
close to the epicenter (Fig. 5.3.2). It has been difficult to differentiate the tsunami signal from
the noise produced by the earthquake, however it is notable that the perturbation exist after the
earthquake time, changing the initial conditions in the calm water level. This information led

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-48
INOCAR to issue a Tsunami Warning over the entire Ecuadorian coast which remained in
effect for 4 hours, considering potential threat for the shores of the nearby Galapagos Islands.

Figure 5.3.2. Tsunami water level during the main earthquake registered at the local INOCAR-DART
buoy which is close to the epicenter (INOCAR, 2016).

Nonetheless, several persons from the coastal communities reported changes in the
seawater level and also strong rifts currents, although the amplitude of the tsunamis did not
reach any significant high water level. Fortunately for the Ecuadorian coastal communities, the
tsunami impact occurred at low tide. This fact is apparently the reason that no inundation
occurred and considerable physical effects at the coastlines have been absent. However, this
case could have been catastrophic, as some authorities declared soon after the main event no
tsunami warning due to the fact that the seismic event has had an epicenter in the continental
side and has not been a marine quake (Toulkeridis et al., 2016).

GEER-ATC Muisne, Ecuador, Earthquake 5-Ground Motions


Report Version 1 5-49
Blank Page
CHAPTER 6
Reconnaissance Methods

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
6.1 Geotechnical Exploration Methods
INTRODUCTION

During reconnaissance, the GEER-ATC team conducted geotechnical investigations and


tests at characteristic locations throughout the area affected by the Mw7.8 Ecuador earthquake
for site characterization. The tests continued after the team left Ecuador by our local GEER
engineers. This section provides some background on the methods used for the geotechnical
investigations including the Standard Penetration Test (SPT), Cone Penetration Test (CPT),
and geophysical in-situ tests using the Downhole Seismic (DS), Multi-Channel Analysis of
Surface Waves (MASW), Microtremor Array Measurements (MAM), and the Horizontal-to-
Vertical Spectral Ratio (HVSR) methods.

STANDARD PENETRATION TEST (SPT)

The Standard Penetration Test (SPT) is the most utilized geotechnical field test in Ecuador,
following ASTMs D1586 and D6066. The SPT is generally used to characterize granular soils
and also stiff clays, where Shelby tube sampling is difficult. During the test, repetitive blows
are applied on an anvil by a free falling hammer weighing 63.5 kg from a height of 76 cm. The
forces on the anvil are transferred to a split-spoon soil sampler using a series of metallic rods.
Each weight drop transfers a potential theoretical energy (E*) of 475 J to the anvil.

Figure 6.1.1. Standard Penetration Test: Driving split-barrel sampler (Mayne et al., 2002).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-1
The number of blows needed to penetrate the sampler for the last 300 mm of a total of a
450 mm depth is the SPT N-value (Fig. 6.1.1). The N-value achieved is inversely proportional
to the energy transferred to the sampler (Schmertmann & Palacios, 1979). It is important to
measure the applied energy, since part of it is lost at various mechanical components in the rod
assembly. For field-measured SPT Nfield-values to be applied in geotechnical engineering
problems, adjustments should be made for effects of hammer energy, overburden pressure and,
various other factors, to normalize the Nfield to N60 or N1,60 (McGregor & Duncan, 1998), where:
 N60 is the blow count corrected to 60% of the theoretical free-fall hammer energy
N60 = Nfield  CE  (CR  CB  CS)
C are corrections of: CE – energy; CR – rod length; CB – borehole diameter; Cs – liner.
 N1,60 normalizes N60 to an effective overburden pressure of 100 kPa
N1,60 = CN  N60 , CN is the overburden correction factor.

CONE PENETRATION TEST

The static Cone Penetration Test (CPT) advances a cone penetrometer through the soil with
a continuous push at a speed of 2 cm/s (ASTM D3441). The penetrometer has strain gauge
cells that measure simultaneously the forces on the cone and friction sleeve (Begemann, 1963;
De Ruiter, 1972, 1981). The cone apex has a 60º angle and base area of ~10 cm2, a 13.25-cm
long sleeve and friction surface of about 150 cm2. The setup of Fig. 6.1.2 is used in Ecuador.

Figure 6.1.2. Cone Penetration Test: Procedure and components (Mayne et al., 2002).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-2
The hydraulic penetration mechanism can apply from 2.5 to 20 tons of axial force. The
resistances at the tip and side are measured by load cells or a similar apparatus. The resistance
qc is calculated by dividing the force acting on the cone Qc, by the projected area of the cone
Ac (qc = Qc / Ac). The friction ratio (Rf) is the sleeve friction (fs) expressed as a percentage of
the cone resistance (qc), both measured at the same depth (Rf = [ fs / qc ]  100%).

SURFACE WAVE METHODS

Surface wave methods use the dispersive nature of Rayleigh waves to determine changes
in subsurface stiffness, and are broadly split in: (i) active and (ii) passive source methods.
Active source methods use a linear array of sensors to measure the phase velocity of waves
emanating from a known source (typically in-line with the array) and propagating past the
receivers. Passive methods traditionally employ 2D sensor arrays to measure the phase velocity
of microtremors (environmental noise) from unknown sources (Tokimatsu et al., 1992). 2D
rather than linear arrays are needed to determine the propagation direction of waves from
unknown sources. In a dispersive material, knowledge of the wave propagation direction and
the true propagation velocity are mutually dependent: one cannot be calculated without the
other (Zywicki, 2007). Passive methods using linear arrays of receivers (Louie, 2001; Park &
Miller, 2008) have been introduced to simplify sensor layout and speed up field data collection.

Figure 6.1.3. Surface wave method: from field acquisition to final Vs profile (C. Wood, U. Arkansas).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-3
Active methods are more capable of resolving shorter wavelengths (higher frequencies),
while passive ones can resolve longer wavelengths due to the low-frequency energy in
microtremors. Following the experimental curve, an inversion process is used to develop the
shear wave velocity, Vs , profile. A numerical solution propagates Rayleigh-type waves over a
layered halfspace (Foti et al., 2014). Iterations are made until the theoretical and experimental
dispersion curves match (Fig. 6.1.3). There is an ongoing effort to combine both methods and
resolve a wider range of wavelengths (Yoon & Rix, 2004; Park et al., 2005; Wood et al., 2014).

MASW
The Multi-Channel Analysis of Surface Waves (MASW, Fig. 6.1.4) active source method
(Park et al., 1999) uses a linear array of 24-48 equally spaced receivers to measure phase
velocities (Zywicki, 1999). An advantage is the ability of MASW to separate higher- from
fundamental-mode surface waves in the experimental analysis, which speeds up the data
processing. Several dispersion analysis techniques are available to process the raw recorded
signals (f-k, f-p, Park transform, beamformer) with the beamformer (Zywicki, 1999) providing
the highest resolution according to Tran & Hiltunen (2008). A fundamental and higher mode
inversion analysis can be used to match the experimental data and obtain a V s profile.

Figure 6.1.4. MASW process from field acquisition to final Vs profile (C. Wood, U. Arkansas).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-4
MAM
The Microtremor Array Measurements (MAM) method was first developed in Japan, using
passive energy to assess the Vs profile with 2D receiver arrays (Aki, 1957; Tokimatsu, 1997).
The arrays can be set using 5-24 receivers in various patterns (circular, nested triangular, L
shapes) and resolve both the surface wave propagation direction and true velocity, regardless
of the propagation direction. Microtremor noise recordings can be taken from 10 min to several
hours to improve spectral estimates (Wood et al., 2014). Three methods are available to process
the data (Wathelet et al., 2008): (i) frequency-wave number (f-k); (ii) high-resolution
frequency-wavenumber (HRFK); (iii) modified spatial auto-correlation (MSPAC). In
determining the best spectral estimate, consideration should be given in the results from all
three methods. Using the spectral estimates, the data typically fit with a multi-mode algorithm
to obtain the Vs profile. The MAM process is shown on Fig. 6.1.5.

Figure 6.1.5. MAM process from field acquisition to final Vs profile (C. Wood, U. Arkansas).

HVSR
In addition to traditional site characterization methods, the horizontal-to-vertical spectral
ratio (HVSR) method (also called Nakamura technique) can be used to determine a site’s

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-5
fundamental vibration frequency that is used to constrain the Vs profile development, and
provide additional information on the location of bedrock or other strong subsurface impedance
contrasts (SESAME, 2004). The method, first introduced by Nogoshi and Igarashi (1971), can
be applied well in microzonation studies and mapping bedrock across large areas. It utilizes a
single 3-component, usually broadband, sensor to record microtremors from the surface. For
processing, records are transformed in the frequency domain, smoothed, and the geometric
mean of the horizontal components are divided by the vertical component. If there is a strong
velocity contrast in the subsurface (e.g., an impedance ratio greater than ~2), a peak will form
in the frequency-spectral ratio domain that will be equal to the fundamental site period.

The described surface wave methods can be used individually to develop a site-specific Vs
profile. To reduce uncertainties and better estimate the Vs over a broad depth range, multiple
methods should be considered when possible. Combining active source methods, such as
MASW, with passive ones, such as MAM, along with HVSR provides additional constrain to
the inversion problem needed to develop the Vs profile (Wood et al., 2014).

TECHNIQUES AND EQUIPMENT USED IN THE ECUADOR MISSION

The reconnaissance team performed MASW testing using a linear array of 24, 4.5-Hz
vertical geophones at a 2-m equal spacing. A typical MASW field setup is shown on Fig. 6.1.6.

Figure 6.1.6. Reconnaissance MASW testing setup with GEER team members K. Rollins, A. Caicedo,
C. Wood working on-site (GPS coordinates: 0o 42' 1.6'' S, 79 o 59' 21.7''W).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-6
Rayleigh-type waves were generated using a 5.4-kg sledgehammer, with sources placed 5
to 40 m from the first geophone. Multiple source offsets applied to insure high quality data,
allow estimation of uncertainties, and ensure that near-field effects did not corrupt the data. A
geometrics Geode seismograph recorded the wave propagation with at least 5 sledgehammer
blows stacked to increase the signal-to-noise ratio. The active source data were analyzed using
the FDBF method (Zywicki, 1999), coupled with the multiple source-offset technique to
identify near-field contamination and quantify dispersion uncertainties (Cox & Wood, 2011).
Dispersion data were generated from each source offset location (5 to 40 m), and the maximum
spectral peak for each frequency was selected automatically using a Matlab code. A composite
dispersion curve was compiled from the individual data of each offset. After eliminating near-
field data, the composite experimental curve was equally divided into 100 frequency bins from
1 to 125 Hz on a log scale. For each bin, the mean phase velocity and standard deviation 
were calculated, resulting in a dispersion curve with associated uncertainty for each frequency.

2D MAM was conducted using an L-shaped array of 24- and 4.5-Hz vertical geophones
spaced at ~3.5 m, depending on the available space at each particular site (Fig. 6.1.7). Passive
energy from urban noise or environmental energy from wind or other sources was utilized. The
surface wave propagation was recorded using a geometrics Geode seismograph.

Figure 6.1.7. Reconnaissance active MAM field testing setup.

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-7
Records of 20- and 30-s were taken at each site for a total time of 10 min. The passive L-
array data (i.e., MAM) were analyzed using the 2D high resolution HRFK method (Capon,
1969) and the 2D f-k method incorporated in the Geopsy software (www.geopsy.org). For the
two methods, vertical vibrations from the L-arrays were used to compute the Rayleigh wave
dispersion data. The uncertainty was estimated by processing the records in 30-s blocks. The
phase velocities from each block were used to obtain a mean and  at each frequency f. The
Rayleigh wave velocity at each f was computed by selecting the x and y wavenumber pair that
produced the peak output power of the wavenumber spectrum. The composite experimental
curve was divided equally into 100 frequency bins from 1 to 40 Hz on a log scale. Similarly as
with the active source, the phase velocity mean and  were calculated for each bin.

HVSR data was collected at selected sites using Nanometrics Trillium Compact boardband
3-component seismometers (Fig. 6.1.8). The sensor has a flat response from 100 Hz to 20 s.
Waveforms were recorded using a Nanometrics Centuar 3-component digitizer and passive
energy was recorded for 10 to 60 min. Geopsy data analysis was performed using squared
average of the N-S and E-W horizontal components. The time records were divided into 60-s
blocks for processing. After rejecting time windows with excessive noise, the remaining
windows were used to create a spectral average for each array. The data were processed in
general accordance with the SESAME (2004) guidelines.

Figure 6.1.8. Reconnaissance HVSR setup and components.

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-8
Once the Rayleigh wave dispersion estimates were obtained and the H/V peak was
determined (if observed), inversion was performed using either the Geospy mulitmodal joint
or the WinSASW effective mode approach. The latter was used only at sites with high velocity
at the surface layer where Geopsy had problems with the inversions calcuations. For the
Geospy inversions, the Rayleigh dispersion data and H/V peak were jointly inverted. Layer
interfaces were only loosely constrained and if layering was indicated from invasive tests, they
were used to constrain the layering. The Vs values of each layer were constrained within
representative Vs geotechnical values, defined as a function of material type and confining
pressure. The Poisson’s ratio  was varied from 0.25 to 0.50, except if the water table was
known, in which case the P-wave velocity Vp was taken as 1,500 m/s and the layer density 
was taken as 2 Mg/m3 regardless of depth or suspected material type. Multiple inversion
analysis were conducted for each site to assess the variability on the Vs profile. The
neighborhood algorithm in Geopsy (Wathelet, 2008) was run with 500,000 models searched in
each analysis. The relative quality of fit was quantified by the minimum misfit, which
compares the theoretical data to the experimental data in terms of the collective squared error.

For dispersion data inverted using WinSASW, Rayleigh disperison data were visually fit
using the 3D theoretical solution by Joh (1996). The solution is approappratie for sites with
inverse profiles which could generate higher mode disperison data and result in
smearing/superposition of modes that can exist in MASW dispersion data at very high or low
freqencies due to lack of spatial resolution. The  was set to 0.3, except where the water table
was known in which case Vp was taken as 1,500 m/s. The density of each layer was taken as 2
Mg/m3 regardless of depth or suspected material type. The Vs profiles from the inversions for
each site were limited to the maximum experimental wavelength divided by two (max / 2).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-9
6.2 Drones
Drones were used to obtain geotagged photos of the affected buildings and areas, as well
as for 3D modeling of buildings and NADIR imaging, using Pix4D for map processing. In the
cities of Manta and Portoviejo, the reconnaissance team used the two types of drones shown
on Fig. 6.2.1: Phantom 4 and Inspire 1 Pro (DJI, 2016a, b). These quad copters record high-
resolution geotagged videos and images that proved invaluable for damage assessment at close
proximity. Using Pix4D mapping software, we reconstructed in 3D some of the areas with
damaged buildings (Fig. 6.2.2). NADIR images were used to create orthographic maps for the
cities of Manta and Portoviejo (Fig. 6.2.3).

Figure 6.2.1. Reconnaissance drones: Phantom 4 (left, DJI, 2016a); Inspire 1 Pro (right, DJI, 2016b).

Figure 6.2.2. 3D mapping examples using Pix4D processing software: (a) Manta.

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-10
(b)

Figure 6.2.2. 3D mapping examples using Pix4D processing software: (b) Portovejo.

Examples of 3D modeling are shown on Fig. 6.2.4 for: (a) the IEES Hospital in Manta
(altizure.com/project/573db054d75edace1f8c7cc3/model), (b) a macro-landslide near San
Isidro (altizure.com/project/573f8f29d75edace1f8ceaaa/model), and (c) a church in
Montecristi, Manabi (altizure.com/project/573e57bbd75edace1f8c9e22/model). The models
were developed by team member Luis Rodriguez More details about the employed
methodologies can be found in Section 6.5.

Figure 6.2.3. Orthomap example taken at an altitude of 37 m.

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-11
(a)

(b)

(c)

Figure 6.2.4. 3D modeling developed by team member L. Rodriguez for: (a) IEES Manta Hospital
(~GPS 0° 57' 18.2" S, 80° 43' 27.0" W); (b) landslide near San Isidro (~GPS 0° 25' 0.6" S, 0° 15' 40.1"
W); and (c) church in Montecristi, Manabi (~GPS 1° 3' 0.0'' S, 80° 40' 0.0'' W).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-12
6.3 Helicopter Flyovers
The Ecuadorian Army Corps of Engineers under the leadership of General Pedro Mosqueda
and coordination by Leutenant Colonel and reconnaisance team member Enrique Morales
organized a flyover with an air force helicopter (Figs. 6.3.1, 6.3.3). The flight was
approximately 4-hours long and scanned most of the damaged areas, as well as three dams, the
coastline and Highway E15. The area covered ranged in latitude between Portoviejo and
Pedernales, with the actual reconnaisance routes shown on Fig. 6.3.2.

Figure 6.3.1. Reconnaissance team members before boarding the Ecuadorian Army helicopter at the
Portovejo army base. From left to right: R. Gilsanz, V. Diaz, D. Alzamora, S. Nikolaou, C. Wood, X.
Vera, K. Rollins, A. Caicedo, R. Luque, G. Lyvers, (GPS 1° 3' 7"S, 80° 28' 30"W).

Figure 6.3.2. GPS track of the reconnaissance helicopter flight provided by the Ecuadorian Army.

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-13
Figure 6.3.3. Ecuadorian air force helicopter (bottom), pilot (upper left corner) and crew officers
(middle left and 3rd right row). General Pedro Mosqueda of the EC Army Corps of Engineers is
discussing the rules of the flight and the routes with reconnaissance team members (2nd right row).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-14
The flight offered important observations and input for subsequent reconnaissance target
decisions on the ground. For example, as the helicopter was approaching the epicentral area,
satisfactory performance of steep slopes adjacent to the highways was observed and landslides
adjacent to the highways were observed (Fig. 6.3.4). While flying over the Chone Dam (Fig.
6.3.5), the observed slope damage led us to return and take shear wave velocity measurements.
Liquefaction and lateral spreading was observed along the rivers that serve as discharge of
dams (Poza Honda, La Esperanza, Chone) and at the port of Manta (Fig. 6.3.6). Liquefaction
evidence and failures within zones hard to be reached by foot or car were easily detected from
the helicopter (e.g., camaroneras shrimp farms observations on Fig. 6.3.7).

Figure 6.3.4. Highway E15 flyover observations: stable steep slopes (left) and landslide (right). (GPS
left: 0° 7' 15.86" S, 80° 13' 15.28" W; GPS right: 0°33' 5"S, 80°25'42"W).

Figure 6.3.5. Chone Dam flyover observations (GPS 0° 51' 58.3"S, 80° 6' 37.6" W).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-15
Figure 6.3.6. Manta port from flyover at an altitude of 35.7 m (GPS 0° 56' 29" S, 80° 43' 20" W).

Figure 6.3.7. Camaroneras farms from an altitude of 92.9 m (top, GPS 0° 37' 38.4" S, 80° 22' 47.9"W).
Liquefaction evidence near Bahia and S. Vicente (bot. left) and slope failure near Pedernales (bot. right).

GEER-ATC Muisne, Ecuador, Earthquake 6-Geotechnical Investigations


Report Version 1 6-16
6.4 Security Cameras
INTEGRATED SECURITY SERVICE (SIS) ECU 911

Ecuador utilizes an Integrated Security System (Servicio Integrado de Seguridad), SIS


called ECU 911 (Fig. 6.4.1). Every organization linked to the network has access through one
single number: 911. This entity is in charge of coordinating the intervention of relevant
institutions and mobilizing resources in the case of accidents or disasters. The ECU 911 system
includes the following organizations: National Police, Armed Forces, Fire Brigade, National
Transit Commission, Ministry of Public Health, Ecuadorian Institute of National Security, Risk
Management Agency, Ecuadorian Red Cross and other local agencies responsible for
emergency care.

Figure 6.4.1. Integrated Security System ECU 911 operation center to monitor emergencies in Ecuador.

The SIS uses a modern technological platform integrated with policies, regulations and
processes, including: video surveillance services, panic buttons, community alarms, reception
and dispatch of emergency teams through calls with the coordination of public institutions,
agencies or entities responsible to give answers to citizens in these situations (Fig. 6.4.2).

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Methods


Report Version 1 6-17
Figure 6.4.2. Services that ECU 911 provides to citizens include (clockwise from top left): video
surveillance; emergency phone lines; community outreach; and institutional coordination.

ECU 911 has a total of 16 centers, 7 of which are distributed amongst the different zones
that the country has been divided into for emergency purposes, and 9 of which are local centers.
During the main earthquake of April 16th, ECU 911 had 1,115 surveillance cameras for active
in the six affected provinces. The video feed kept a record of the places were rescue operations
were conducted. The system has the capacity of monitoring any irregularities to avoid further
complications such as looting and other illegal actions. In the following sections, images
captured on the day of the earthquake are presented for selected affected cities.

GUAYAQUIL
Images from the city of Guayaquil are presented in Figs. 6.4.3 to 6.4.5.

(a) (b)
Figure 6.4.3. Images from camera GYE 61F1 near downtown Guayaquil (9 de Octubre and Padre
Solano Streets): (a) before, and (b) during the main seismic event.

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Methods


Report Version 1 6-18
Power outages caused by the earthquake were captured as shown on Figs. 6.4.3 and 6.4.5.
A structural collapse and the street after the event were captured as shown on Fig. 6.4.4.

Figure 6.4.4. Camera GYE 157 D images (Chile & 9 de Octubre Sts.), before, during, after the event.

Figure 6.4.5. Images from camera GYE 146D in downtown Guayaquil (9 de Octubre & P. Carbo St.).

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Methods


Report Version 1 6-19
MANTA
Images obtained in Manta are presented in Figs. 6.4.6 and 6.4.7 with power outages.

Figure 6.4.6. Camera Manta23 images from the avenue before (upper left) and after the event.

Figure 6.4.7. Camera Manta31 images showing Barrio Altamira, before (upper left) and after the event.

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Methods


Report Version 1 6-20
SAMBORONDÓN
Samborondon city images are shown on Fig. 6.4.8 with power outage in the bottom image.

Figure 6.4.8. Images from SAM07F at la Puntilla (3.3-km), before (top) and during the event (bottom).

PORTOVIEJO

Figure 6.4.9. Camera Porto09 images captured Manabí and Atahualpa St. near Imperial Hotel before
(top), during (middle) and after the seismic event (bottom).

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Methods


Report Version 1 6-21
PRIVATE CAMERA SECURITY VIDEOS

Security videos were also released from offices and other facilities. The example of Fig.
6.4.10 shows images from an office camera in Canoe (youtube.com/watch?v=Nz_05mT-Me0).

Figure 6.4.10. Camera images from an office in Canoe (youtube.com/watch?v=Nz_05mT-Me0).

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Methods


Report Version 1 6-22
6.5 Reconnaissance Tools
The reconnaissance team took numerous photographs and videos with conventional
cameras and their smart phones (Fig. 6.5.1). As per GEER procedures, our team geo-tagged
every photo, with the GPS locators in smartphones and digital cameras proven to be accurate
means to identify (e.g., Fig. 6.5.2 shows the located Manta strong ground motion station).

Figure 6.5.1. GEER-ATC members taking reconnaissance photos: V. Diaz and G. Lyvers, Portovejo
(left, 1o3'16.5''N, GPS: 80o26'42.3''E); C. Wood and others, Manta (top right, GPS 0o56'28.3''N,
80o44'4.1''E); R. Luque, helicopter 83m above Los Cuyeyes (bot. right, GPS 1o6'57.7''N, 80o11' 29.5''E).

Handheld GPS watches (Fig. 6.5.3) were used mostly to confirm locations among team
members and to geo-tag photos from conventional cameras, whose batteries would run out
quickly if they had GPS options. In addition to geo-tagging, our team had to track
reconnaissance routes. For these purposes, we were equipped with handheld GPS tracking units
that continuously record the owner’s route. These devices (e.g., GPSMAP®64 by Gamin, Fig.
6.5.3) are rugged and water resistant, have large, 6.6-cm, sunlight-readable color screen,
preloaded basemaps, and strong reception using a quad helix antenna. External photos could
be transferred to the device’s database with downloads of information and at the end of the
day. Examples of routes of two GEER-ATC teams on April 29th are shown on Figs. 6.5.4.

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Tools


Report Version 1 6-23
Figure 6.5.2. Manta strong motion accelerometer station with coordinates identified by smartphone.
On the photo: team member G. Ponte Vasquez (GPS: 0o 56' 28'' S, 80o 44' 3'' W, Elev. 33 m).

Figure 6.5.3. Handheld GPS tracking device types used in reconnaissance (Garmin GPSMAP®64
tracker - middle; Forerunner 305 Running watch - top right; Foretrex 401 Hiking watch - bottom right).
One member of each team deployed had a tracking device (see C. Wood on the left, Manta, GPS 0o 56'
29.9'' N, 80o 43' 57.1'' E). Additional devices were coordinated for accuracy at the start of each mission.

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Tools


Report Version 1 6-24
Figure 6.5.4. GEER-ATC reconnaissance survey routes on April 29, 2016 in Pedernales, Manabí: team
1 (left) and team 2 (right). Base Image by Google Earth (©2016).

In addition to GPS positioning, cameras, and video recording equipment, the team took
basic measurements for geotechnical and structural observations of interest. Such
measurements are not intended to replace detailed topographic mapping using surveying
equipment, but may provide indicative, expedited measurements of post-disaster patterns.
Basic equipment includes tape and laser-based tools that can accurately measure distances,
lengths, offsets, and crack widths or tilts (Fig. 6.5.5). Often, visual inspections and order of
magnitude of damage was observed by comparing to an object or person (Fig. 6.5.6).

Measurements may be performed along specific lines like the lateral and vertical
settlements observed along the Manta port interface between the quay wall and the sidewalk
inland that were measured conventionally in equal spaces along the settlement line, as well as
with visual estimate in hard to reach places as shown on Fig. 6.5.5. This information is
necessary to get an estimate of the direction and magnitude of movements and their spatial
distribution, as described in Chapter 6.

Smartphone levelling apps were used and gave fast estimates of offsets from the vertical
position of post-earthquake rotated walls or lamp posts (Fig. 6.5.7). An example of
measurements of soil deformations and associated pavement failures between monuments at
the Portovejo cemetery is shown in Figure 6.5.8.

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Tools


Report Version 1 6-25
Figure 6.5.5. Manta port deformations along the interface between quay wall and sidewalk (GPS 0o 56'
13.2'' N, 80o 43' 27.4'' E). Measurements using metallic tape (bot. left) by A. Caicedo (top, right) and
visual evaluation of deformations in areas hard to reach by D. Alzamora and A. Zekkos (bottom, right).
Laser-based equipment (center) were used to measure distances (left) and tilts (right).

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Tools


Report Version 1 6-26
Figure 6.5.6. Manta port wall fissure (left, K. Rollins, GPS 0o56'29.6''N, 80o43'27.4''E); Portovejo
cemetery (top right, A. Caicedo, GPS 1o3'14.6''N, 80o26'52.1''S); Manta R. Zambrano Hospital bottles
(bottom, G. Lyvers, GPS 0o57'15.4''N, 80o44'32.8''E).

Figure 6.5.7. Manta port parking lot wall (GPS 0o 56' 24.3'' N, 80o 43' 32.7'' E) tilting measurements
using rigid tape (A. Caicedo, left) and smartphone app (A. Zekkos and K. Rollins, right).

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Tools


Report Version 1 6-27
Figure 6.5.8. Soil deformations and pavement failures between monuments at the Portovejo cemetery
measured by team member G. Lyvers (GPS 1o 3' 14.4'' N, 80o 26' 49.2'' E).

GEER-ATC Muisne, Ecuador, Earthquake 6-Reconnaissance Tools


Report Version 1 6-28
Blank Page
Blank Page
CHAPTER 7
Geotechnical Observations

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
7.1 Geologic Setting
Ecuador rises eastward from the Pacific Ocean to form two imposing Andean cordilleras
(mountain ranges or chains) before descending into the sedimentary deposits of the Amazon
rainforest and includes the Galapagos Islands, located about 1,500 km west of the mainland.
Covering 620 km of the Andes, with a surface area of approximately 284,000 km2, the country
is part of the Circum-Pacific belt, north of the Huancabamba oroclinal deflection that marks a
geological and structural boundary. Ecuador’s geography and geology has been documented
as far back as 1892 when Teodoro Wolf published the book “Geografía y Geología del
Ecuador.” Since then (e.g., Holloway, 1932), Ecuador’s geological and mineralogical makeup
has been documented in several publications from the 60’s to the 90’s including international
collaborations with other countries. Detailed information was produced by PRODEMINCA,
the environmental control and mining development supported financially from 1997-99 by the
World Bank and the governments of the United Kingdom, Sweden, and Ecuador.

Figure 7.1.1. Geo-structural domains of Ecuador (Litherland and Zamora, 1991; Bolaños, 2010).

Proceeding west to east, Ecuador covers six geo-structural divisions as shown on Fig. 7.1.1
(Litherland and Zamora, 1991; Bolaños, 2010); namely the: (i) Fore Arc Basin of the coast, (ii)
Western Cordillera, (iii) Interandean Graven, (iv) Real or Central Cordillera, (v) Eastern Sub-
Andean zone, and (vi) Back Arc Basin of Iquitos. These divisions are subdivided into sub-
regions, each having their own characteristics such as habitat, climate, and soil types, modified
by natural factors such as weathering, erosion, and biochemical activity (Vera, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-1
A geologic map of Ecuador, N. Peru and S. Colombia, is shown on Fig. 7.1.2, compiled by
Spikings et al. (2000) from data in Baldock (1982a, b), Litherland & Zamora (1991), Toussaint
& Restrepo (1994), and Litherland et al. (1993). Early Cretaceous to present day convergence
between the Pacific oceanic and South American continental plates has created distinguishable
tectono-stratigraphic regions, such as the Cordillera Real (Litherland et al., 1994), the Sub-
Andean Zone (SAZ) and Oriente region. Off-shore bathymetry identified the Carnegie Ridge
and the Peru-Ecuador-Colombia trench. The three traverses across the Eastern Cordillera are
labeled as: (a) Julio Andrade - La Bonita, (b) Papallacta - Baeza, and (c) Banos - Puyo. Also
shown are the major faults Banos (BF), Cosanga (CF), lngapirca (IN), Llanganates (LF), and
Peltetec (PF).

Figure 7.1.2. Geologic map of Ecuador, northern Peru, and southern Colombia (after Spikings et al.
2000) with fault notations; Banos (BF), Cosanga (CF), lngapirca (IN), Llanganates (LF), Peltetec (PF).

The Andes orogenic belt was composed by several Mesozoic-Cenozoic events. In Early
Cretaceous ~140-120 Ma (Feininger & Bristow, 1980), the Cordillera Real formed the
easternmost mountain chain and originated deposition of terranes, known as the Peltetec Event
(Litherland et al., 1994). Since Peltetec, growth of the Cordillera Real has exposed schist faces
and lower-grade metamorphic rocks (Davila & Eguez, 1990). The simplified bedrock map of
Ecuador of Fig. 7.1.3 (A. Alden, USGS, 1997) shows sedimentary rock formations at the
coastal west and igneous and metamorphic formations at the mountainous areas inland.
GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations
Report Version 1 7-2
Figure 7.1.3. Bedrock Geology. Sedimentary rock formations on the coastal west and igneous and
metamorphic rock formations at the mountainous areas inland (from A. Alden, USGS, 1997).

The bedrock formations have been dated using the metamorphic micas radiometric
Potassium-Argon (K-Ar) method that measures radioactive decay of an isotope of potassium
into argon. The K-Ar ages are shown on Fig. 7.1.4 range from Jurassic to Early Cretaceous
metamorphic rocks and Tertiary plutons of Cordillera Real. Most rocks appear to be less than
100 Ma old, with pre-Cretaceous metamorphic ones are from 90 to 50 Ma, suggesting that they
have been reset to varying degrees (Litherland et al., 1994).

Figure 7.1.4. Bedrock formation K-Ar ages from Jurassic to Early Cretaceous metamorphic rocks and
Tertiary plutons of the Cordillera Real (Litherland et al., 1994).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-3
The Ecuadorian continental margin has been active since the late mid-Jurassic about 160
Ma (Fig. 7.1.5A), as discussed in Spikings et al. (2000); Pilger (1983); Pardo-Casas & Molnar
(1987); Gallagher (1989); and Jaillard et al. (1990, 1995). Within the Sub-Andean Zone,
tectonic uplift is evident and a major fault separates this zone from flat-lying Late Cretaceous
- Recent Sedimentary rocks of the Foreland or "Oriente" Basin, as South America split from
Africa in mid-Jurassic and started moving west (Brookfield et al., 2009). Rifting began in mid-
Jurassic, followed by the creation of a marginal basin spreading ridge in the Cretaceous (Lebrat
et al., 1986). Deformation and uplift in the Andes to the west, beginning in early Cenozoic led
to deposition of thick coarse successions (Fig. 7.1.5B), and progressive deformation of both
Mesozoic and Cenozoic sediments across the basin (Feininger and Bristow, 1980).

Figure 7.1.5. Mesozoic evolution of Ecuador (Kennerley, 1980; Brookfield et al., 2009).

The coastal Cretaceous to Cenozoic basin underlain by aloctonous basaltic ocean crust
(Ocean Piñon Terrane) is the flat zone that was hit the hardest in the 2016 earthquake (Fig.
7.1.2). The most recent Cretaceous (Fig. 7.1.5C) seismotectonic and volcanic events are
described in detail in Chapter 4. Geotechnical information on the overburden soils of the
affected areas from the 2016 earthquake are provided in the next section of this chapter.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-4
7.2 Site Characterization Data
INTRODUCTION

Geotechnical and geophysical subsurface investigation data include available information


pre-earthquake from projects in design or in construction phases, and from post-earthquake
reconnaissance GEER investigations. Figure 7.2.1 shows the sites with available geotechnical
data that will be presented in this section, with details on the locations and types of geotechnical
testing performed. Details on the methods used in the subsurface investigations are provided
in Chapter 6.

PRIOR TO EARTHQUAKE

RECONNAISSANCE
AFTER EARTHQUAKE

100 mi
160 km

Figure 7.2.1. Site locations with pre- and post-earthquake geotechnical data (Google Maps ©2016).
Details for each site are presented on Table 7.2.1.

Pre-earthquake subsurface information was provided to our team by the Ecuadorian


Ministry of Public Works and Transportation and directly from local engineering and
construction firms. As per local geotechnical practice standards, the subsurface investigations
in the available geotechnical reports include conventional borings with Standard Penetration
Tests (SPT), laboratory sample re-classification and testing of Atterberg limits, water content,
unconfined compression, and consolidation. In few cases, California Bearing Ratio (CBR)
penetration (road projects) and in-situ shear wave velocity Vs tests were available.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-5
Post-earthquake field tests were performed during and after reconnaissance, focusing on
sites of interest (such as locations of strong motion stations or liquefied areas). The
reconnaissance field geophysical work was led by GEER member Professor C. Wood and
supported and executed in collaboration with the local companies Geoestudios and Subterra.
The geophysical reconnaissance testing included in-situ Vs measurements using the techniques
MAM (Microtremor Array Measurements) and MASW (Multichannel Analysis of Surface
Waves) and Horizontal-to-Vertical Spectral Ratio (HVSR) tests, combined with SPT and sCPT
with pore-pressure measurements (sCPTu) in nearby locations. Note that a total of 7 HVSR
tests (HV1 to HV7) were performed in the vicinity of site 20 in Table 7.2.1 (HV1 in Manta).

Table 7.2.1. Locations with geotechnical information pre- and post-earthquake.


SITE IDENTIFICATION GEOGRAPHIC COORDINATES GEOTECHNICAL TESTING

No. Source Name Station City Province Latitude Longitude SPT CPT MAM MASW HVSR

PRE-EARTHQUAKE

1 Manta Port Manta Manabi 0°55'58.73"S 80°43'19.15"W 


Pre-EQ Project Data

2 Mejia Bridge Portoviejo-Crucita Manabi 0°59'24.00"S 80°28'12.00"W 

3 Los Caras Bridge Bahia de Caraquez Manabi 0°36'36.06"S 80°25'24.35"W 

4 Los Chacras Bridge Pechicho Manabi 0°58'35.69"S 80°25'21.72"W 

5 Briceño Bridge Boca de Briceño Manabi 0°31'18.11"S 80°26'23.83"W 

6 Chone Dam Chone Manabi 0°42'0.00"S 79°59'20.40"W 

POST-EARTHQUAKE RECONNAISANCE

1 Manta Port Manta Manabi 0°55'58.73"S 80°43'19.15"W     

2 Mejia Bridge Portoviejo-Crucita Manabi 0°59'24.00"S 80°28'12.00"W  

3 Los Caras Bridge Bahia de Caraquez Manabi 0°36'36.06"S 80°25'24.35"W

4 Los Chacras Bridge Pechicho Manabi 0°58'35.69"S 80°25'21.72"W 

5 Briceño Bridge Boca de Briceno Manabi 0°31'18.11"S 80°26'23.83"W

6 Chone Dam Chone Manabi 0°42'0.00"S 79°59'20.40"W   

7 APO1  Portoviejo Manabi 1° 2'11.17"S 80°27'32.88"W   


Post-EQ GEER reconnaissance Investigations

8 AMNT  Jaramijó-Manta Manabi 1° 0'7.70"S 80°33'33.21"W  

9 ACHN  Chone Manabi 0°41'52.80"S 80° 5'2.40"W  

10 APED  Pedernales Manabi 0° 4'4.80"N 80° 3'25.20"W  

11 AMA1  Cojimies-Pedernales Manabi 0°13'35.11"N 79°59'46.82"W  

12 AV18  Esmeraldas Esmeraldas 0°47'53.01"N 79°40'14.22"W  

13 ALIB  La Libertad Santa Elena 2°14'34.80"S 80°50'45.60"W  

14 AMIL  Guayaquil Guayas 2°10'51.60"S 79°31'44.40"W  

15 AGYE  Guayaquil Guayas 2° 3'14.40"S 79°57'7.20"W  

16 GYE1  Guayaquil Guayas 2°15'0.00"S 79°54'36.00"W  

17 GYE2  Guayaquil Guayas 2°11'56.40"S 79°53'52.80"W  

18 Catolica University Guayaquil Guayas 2°10'51.60"S 79°54'14.40"W  

19 Manta SMS Manta Manabi 0°56'27.60"S 80°44'4.56"W   

20 Manta HV1 Manta Manabi 0°57'18.00"S 80°42'32.40"W 

21 IESS Hospital Manta Manabi 0°57'14.40"S 80°43'26.40"W   

22 Mobil Manta Manabi 0°56'52.80"S 80°43'15.60"W 

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-6
PRE-EARTHQUAKE GEOTECHNICAL INFORMATION

Geotechnical data from borings drilled before the main Mw7.8, April 16 2016 earthquake
are available for six major projects located within the coastal province of Manabí consisting of
a port, four bridges, and a dam. Listed in Table 7.2.1, they are: (1) Manta Port; (2) Mejia
Bridge, Portoviejo-Crucita; (3) Los Caras Bridge, Bahia; (4) Las Chacras (Rio Chico) Bridge,
Pechicho; (5) Briceño Bridge; and (6) Chone Dam. The locations are shown in Fig. 7.2.2.
Standard penetration test (SPT) data from borings is available from all six projects. In-situ Vs
measurement data are available from the Manta port project. Post-earthquake, extensive
reconnaissance investigations have been conducted (except for projects 3 and 5) and are
presented later in this chapter.

SITE IDENTIFICATION GEOGRAPHIC COORDINATES

No. Name City Latitude Longitude

1 Manta Port Manta 0°55'58.73"S 80°43'19.15"W

2 Mejia Bridge Portoviejo-Crucita 0°59'24.00"S 80°28'12.00"W

3 Los Caras Bridge Bahia de Caraquez 0°36'36.06"S 80°25'24.35"W

4 Los Chacras Bridge Pechicho 0°58'35.69"S 80°25'21.72"W

5 Briceño Bridge Boca de Briceño 0°31'18.11"S 80°26'23.83"W

6 Chone Dam Chone 0°42'0.00"S 79°59'20.40"W

100 mi

Figure 7.2.2. Locations of 6 pre-earthquake projects with geotechnical data (Google Maps ©2016).

Manta Port, Pre-Earthquake Geotechnical Information

The largest seaport in Ecuador is in Manta, the second largest city of the province of
Manabí. The port, shown in Fig. 7.2.3, is managed by the Autoridad Portuaria Manta (LCA,
2016). It has historic value, since this port was used by Charles Marie de La Condamine when
leading the French mission in Ecuador to measure the location of the equator in 1735. The port
of Manta experienced extensive liquefaction and earthquake-induced settlements during the
main earthquake. Hence, the available geotechnical data can provide useful information on the
pre-earthquake subsurface conditions, including in-situ Vs measurements.
GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations
Report Version 1 7-7
Figure 7.2.3. Manta Port (GPS Coordinates: 0° 57' 35.0" S, 80° 43' 2.0" W). Photo by LCA (2016).

Figure 7.2.4. Manta Port: Pre-earthquake boring locations (Google Maps ©2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-8
Data available from the 2007 expansion of the port (Ripalda, 2007) include logs from 20
borings and 6 test pits, presented in Appendix A (Figs. A1-A26), and 3 in-situ Vs measurements
from Downhole Seismic (DS) tests (Figs. A27-A29). The boring logs present index properties
(water content and Atterberg limits), unit weights, grain sizes, and strength expressed by the
SPT N-value that was recorded for each sample taken. The locations of the borings are shown
on Fig. 7.2.4 and summarized in Table 7.2.2.

Based on the borings and laboratory data, the subsurface profile consists of approximately
10 to 15 m of sand and gravel fill at the top, overlying a dense layer of fine sand and silt with
organic particles extending to a depth of about 45 m below the ground surface. Weathered
siltstone exists below this layer to the maximum explored depth of about 54 m.

Table 7.2.2. Manta Port: Pre-earthquake boring coordinates.


Site Investigation Geographic Coordinates
Test Identification Test Type Latitude Longitude
B110 Boring 0°55'56''S 80°43'13''W
B111 Boring 0°55'50''S 80°43'9''W
B116 Boring 0°55'50''S 80°43'11''W
B117 Boring 0°55'25''S 80°42'58''W
B118 Boring 0°55'40''S 80°42'58''W
B120 Boring 0°55'37''S 80°43'3''W
B123 Boring 0°55'40''S 80°43'9''W
B124A Boring 0°55'36''S 80°43'2''W
B125 Boring 0°55'35''S 80°42'58''W
B130 Boring 0°55'53''S 80°43'16''W
B132 Boring 0°56'0''S 80°43'20''W
B132B Boring 0°55'58''S 80°43'19''W
B134 Boring 0°56'4''S 80°43'25''W
B135A Boring 0°56'4''S 80°43'28''W
B137 Boring 0°56'4''S 80°43'20''W
B138 Boring 0°56'6''S 80°43'23''W
B139 Boring 0°56'11''S 80°43'24''W
B142 Boring 0°56'13''S 80°43'21''W
B143 Boring 0°56'1''S 80°43'14''W
TP-1 Test Pit 0°56'12''S 80°43'31''W
TP-2 Test Pit 0°56'9''S 80°43'30''W
TP-4 Test Pit 0°56'7''S 80°43'28''W
TP-6 Test Pit 0°56'1''S 80°43'19''W
TP-8 Test Pit 0°55'51''S 80°43'12''W
TP-9 Test Pit 0°55'51''S 80°43'12''W

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-9
Mejia Bridge, Pre-Earthquake Geotechnical Information

The Mejia Bridge is located on the road between Portoviejo and Crucita. It is composed of
two bridges alongside, constructed roughly 15 years apart. They span approximately 50 m over
the Portoviejo River (Fig. 7.2.5). The bridges have a combined width of approximately 25 m,
with 4 traffic lanes in total. According to the design drawings, the abutments are supported on
25-m long driven piles. The 20-cm thick concrete bridge deck is supported on steel beams.
Additional information about the Mejia Bridge can be found in Section 7.7.

Figure 7.2.5. View of the Mejia Bridge prior to the earthquake (Google Maps ©2016).

There are two pre-earthquake sets of geotechnical data available for this site, shown on Fig.
7.2.6. The first dates circa 2000 back to the construction of the first bridge, referred to as “old
bridge,” and consists of 54- and 60-m deep borings at each abutment. The second, performed
during the construction of the second bridge, is presented in a geotechnical report with two
boring logs (OB-1 and OB-2) which include SPT N-values, index properties (water content
and Atterberg limits), unit weights, grain sizes, and strength tests using vane shear and torvane.
This boring information is presented in Appendix A, Figs A30-A31 with the coordinates of the
borings shown on Table 7.2.3 and actual locations on Fig. 7.2.6.

Table 7.2.3. Mejia Bridge: Pre-earthquake boring coordinates.


Site Investigation Geographic Coordinates
Test
Test Type Latitude Longitude
Identification
OB-1 Boring 0°59'24''S 80°28'11''W
OB-2 Boring 0°59'23''S 80°28'11''W

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-10
Figure 7.2.6. Mejia Bridge: Geotechnical test locations pre-earthquake (yellow squares) and post-
earthquake. (Google Maps ©2016).

The original set of borings indicate various soil layers that overlay a very dense gravel
layer. At the top of the profile there is a desiccated crust, with shear strength values Su between
60 and 180 kPa. This layer extends to the depth of the water table, approximately 2.9 to 5.4 m
below the surface, with a high-plasticity, soft to medium stiff clay and silt layer extending to
about 20 m with Su between 10 and 80 kPa. Within this layer, there is a 3-m thick sandy clay
layer with average SPT N-value of 15 blows/30cm. From 20 to 24 m depth there is a stiff clay
and silt layer with Su between 80 and 200 kPa. The bottom layer is clay with Su from 60 to 200
kPa (based on N-values) and Plasticity Index PI between 25 and 52 throughout the site.

Los Caras Bridge, Pre-Earthquake Geotechnical Information

The Los Caras Bridge is a critical structure that joins the coastal towns of San Vicente and
Bahía de Caráquez, facilitating transportation in the Manabí province (Fig. 7.2.7). The project
was undertaken by the Army Corps of Engineers of Ecuador and was built in 2010. The bridge
is supported by 48 piers and has a central span between the two abutments with a total length
of 1,985 m. The approaches are via a 208-m access ramp from the Bahía span, and a 607-m
access ramp from the San Vicente span. Additional information about this bridge can be found
in Section 7.7.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-11
Figure 7.2.7. Los Caras Bridge (Flyg London, 2011).

The Los Caras Bridge site is characterized as an estuarine environment formed by alluvial
deposits with depths between 25 and 85 m overlying weathered rocks (Appendix A, Fig. A32).

Las Chacras (Rio Chico) Bridge, Pre-Earthquake Geotechnical Information

Las Chacras Bridge is located north of Portoviejo (GPS 0°41'41.7"S, 79°59'21.9"W, Fig.
7.2.1). The site experienced major embankment failures during the main earthquake with
extensive damage for hundreds of m. A pre-earthquake photo is shown on Fig. 7.2.8. MAM-
MASW tests were performed post-earthquake and will be discussed later in this chapter.

Active MASW Array


Active Source Locations

Meters
0 66

Figure 7.2.8. Las Chacras Bridge: Pre-earthquake photo. Base image: CNES/Astrium (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-12
Two boring logs are available for Las Chacras Bridge, drilled to depths of 32 m and 34 m
below ground surface. The boring logs include SPT N-values, water contents and Atterberg
limits, and are provided in Appendix A, Figs. A33-A36. The coordinates of the borings are
shown on Table 7.2.4 and their actual locations on Fig. 7.2.9.

Table 7.2.4. Las Chacras Bridge: Pre-earthquake boring coordinates.


Site Investigation Geographic Coordinates
Test Identification Test Type Latitude Longitude
P1 Boring 0°58'35''S 80°25'21''W
P2 Boring 0°58'36''S 80°25'23''W

Figure 7.2.9. Las Chacras Bridge: Location of pre-earthquake borings (Google Maps ©2016).

Based on the available data, the subsurface profile consists of approximately 2 to 4 m of


loose silty and clayey sands, followed by a low to medium plasticity layer of stiff silt with
seams of sand and clay extending to a depth of about 10 m below the ground surface. For the
remaining depth explored, a thick layer of dense sandy clay is present.

Briceño Bridge, Pre-Earthquake Geotechnical Information

The Boca de Briceño Bridge is located next to the homonymous town, 6 km south of Canoa.
A new roadway alignment was constructed from 2012 to 2014 (Fig. 7.2.10), to eliminate the
travel via the U-shaped section of the road, without interfering with the scenic views along the
coast. This is part of a project extending from San Vicente to Pedernales which started in 2010.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-13
Figure 7.2.10. Boca de Briceño Bridge aerial photo (GPS: 0o 30' 58.3'' S, 80o 26' 29.1'' W).

Two borings were drilled for the Briceño Bridge project to a depth of 16 m. The logs report
SPT N-values, water contents and Atterberg limits as shown on Figs. A37 and A38 of
Appendix A. The locations of the borings are summarized in Table 7.2.5 and Fig. 7.2.11.

Table 7.2.5. Briceño Bridge: Pre-earthquake boring coordinates.


Site Investigation Geographic Coordinates
Test
Test Type Latitude Longitude
Identification
P1 Boring 0°30'59''S 80°26'30''W
P2 Boring 0°30'56''S 80°26'30''W

Figure 7.2.11. Briceño Bridge: Locations of pre-earthquake borings (Google Maps ©2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-14
Based on the available data, the soil profile consists of 2- to 5-m thick loose silty sand to
sandy silt with the exception of Station 8+425 where the top 2.5 m are a medium stiff clay
layer, followed by interlayers of stiff clay or sandy silt and dense silty or clayey sand, sandy
silt, or very stiff clay to a 5-m depth. Below 5 m, stiff to very stiff clays were encountered.

Ground improvement through Geopiers was used in this project to address the stability and
settlement concerns for the embankment, particularly for seismic conditions.

Chone Dam, Pre-Earthquake Geotechnical Information

Chone Dam is a multipurpose dam currently under construction in the province of Manabí,
located downstream of the confluence of the Platanales and Chone Rivers, about 15 km
upstream from Chone (Fig. 7.2.12, with exact coordinates 0° 41' 59.6" S, 79° 59' 21.5" W).
Construction started in 2011, and the dam is scheduled to be completed and transferred to the
government by the end of 2016. The coordinates for the dam are 0°41'59.6"S, 79°59'21.5"W.
The dam is designed for flood management, irrigation, and water supply to both industrial and
city use. It is owned and operated by the National Secretariat of Water (Secretaría National de
Agua), known as SENAGUA.

Figure 7.2.12. Chone/Rio Grande Dam aerial view (Photo GPS: 0°41'44.1''S, 79°59'30.6''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-15
Ten borings are available to depths up to 50 m. The logs report SPT N-values, water
contents and Atterberg limits as presented in Figs A39-A44 of Appendix A. The boring
locations are shown on Fig. 7.2.13 and the coordinates are presented on Table 7.2.6.

Table 7.2.6. Chone Dam: Pre-earthquake boring coordinates.


Site Investigation Geographic Coordinates
Test
Test Type Latitude Longitude
Identification
P1 Boring 0°42'1''S 79°59'21''W
P2 Boring 0°42'0''S 79°59'21''W
P3 Boring 0°41'59''S 79°59'21''W
P4 Boring 0°41'59''S 79°59'23''W
P5 Boring 0°42'1''S 79°59'18''W
PPMCH-01 Boring 0°41'60''S 79°59'22''W
PPMCH-02 Boring 0°42'1''S 79°59'23''W
PPMCH-03 Boring 0°41'58''S 79°59'21''W
PPMCH-04 Boring 0°41'60''S 79°59'20''W
PPMCH-05 Boring 0°41'59''S 79°59'22''W

Figure 7.2.13. Chone Dam: Locations of pre-earthquake borings (Google Maps ©2016).

Based on the borings, the soils underlying the dam are primarily sands and silts. The
technical report (UCSG, 2015) gives detailed information on the geologic profile of the area,
which is presented in Section 7.8.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-16
POST-EARTHQUAKE GEOTECHNICAL INFORMATION

Geotechnical data were collected during and after the reconnaissance trip of the GEER-
ATC team for 20 sites shown on Fig. 7.2.14 with details on Table 7.2.1.

RECONNAISSANCE
AFTER EARTHQUAKE

100 mi
160 km

Figure 7.2.14. Locations of 20 post-earthquake GEER investigations (Google Maps ©2016).

In summary, the sites and types of tests performed are: (1) Manta Port: borings, CPTu, and
geophysical testing; (2) Mejia Bridge: borings, CPTu; (3) Las Chacras (Rio Chico) Bridge:
geophysical testing; (4) Chone Dam: geophysical testing; (5) Station APO1, Portoviejo:
borings and geophysical testing; (6) Station AMNT, Jaramijo, Manta: geophysical testing; (7)
Station ACHN, Chone, geophysical testing; (8) Station APED, Pedernales: geophysical
testing; (9) Station AMA1, Cojimies, Pedernales: geophysical testing; (10) Station AV18,
Esmeraldas: geophysical testing; (11) Station ALIB, La Libertad: geophysical testing; (12, 13,
14, 15) Stations AMIL, AGYE, GYE1, GYE2, Guayaquil: geophysical testing; (16) Catolica
University, Guayaquil: geophysical testing; (17, 18, 19, 20) SMS, HV, IESS Hospital, Mobil,
Manta: geophysical testing.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-17
Manta Port, Post-Earthquake Geotechnical Information

The Port of Manta is at the northwest part of the city (0° 56' 25.9" S, 80° 43' 31.8" W) along
the Malecon de Manta highway. The port was heavily damaged during the main event and lost
some functionality. The GEER geotechnical post-earthquake investigations are described here.

One boring, P-1 (0° 55' 58.8" S, 80° 43' 19.2" W), was drilled to a depth of 10.2 m with the
water table at a depth of 2.8 m. According to the boring log (Fig. B1, Appendix B, Fig. B1),
below the top 0.2 m of asphaltic concrete, there is a relatively homogeneous light gray poorly
graded sand with silt and presence of shells was sampled throughout the borehole varying
primarily in density. This layer has energy-corrected N60 values from 7 to 75 blows/30cm,
water content, w, between 7 and 24%, and a fines content, FC, between 5 and 11%. Cone
Penetrometer Testing (CPT) was performed next to boring P-1 (0° 56' 2.4" S, 80° 43' 19.2" W)
to the same depth (see data in Appendix B, Fig. B2).

Surface wave measurements were conducted at the parking area and the wharf deck. The
parking area (Figs. 7.2.15 and 7.2.16) experienced liquefaction and lateral spreading during the
main event and sections along the wharf deck settled (Figs. 7.2.19 and 7.2.20).
Active MASW Array
Passive 2D MAM Array
Active Source Locations
Horz to Vert Spectral Ratio

Meters
0 103

Figure 7.2.15. Manta Port parking area before the earthquake with locations of active and passive
surface wave arrays, HVSR. Base image: Digital Globe (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-18
Figure 7.2.16. Manta Port parking area post-earthquake: GEER team member C. Wood at the active
MASW array (GPS: 0° 56' 25.9" S, 80° 43' 31.8" W).

At the parking area, the shear wave velocity, Vs , profile developed and the experimental
dispersion curve are shown on Table 7.2.7 and Fig. 7.2.17. The data indicate a stiff top 2.5-m
thick with Vs of 270 m/s, followed by an approximately 3-m thick soft layer with Vs of 150
m/s and 5.5-m thick stiffer layers. These results suggest that the second layer of Vs = 150 m/s
may had liquefied during the main ground motion shaking.
Overall, the site has an average Vs at the top 30 m, Vs30 of 355 m/s with HVSR results
shown on Fig. 7.2.18. The experimental measurements indicate a HVSR peak at about 5 Hz,
which matches well with the theoretical HVSR peak developed from the in-situ Vs profile.

Table 7.2.7. Manta Port parking area: In-situ measured Vs profile.


Layer Depth to Bottom of Layer Shear Wave
Number Layer Thickness Velocity Vs
# (m) (m) (m/s)
1 2.5 2.5 270
2 5.5 3.0 150
3 11.5 6.0 350
4 25.5 14.0 450
5 60.0 34.5 650
Vs30 (m/s) 355 Vs to depth of 30 m

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-19
700 0

Mean with +/- 1 


600 Theoretical Fit 10

500
Phase Velocity (m/s)

20

Depth (m)
400
30
300

40
200

50
100

0 60
10 100 0 200 400 600
Frequency (Hz) Shear wave Velocity (m/s)

Figure 7.2.17. Manta Port parking area: Median theoretical Rayleigh wave dispersion curve with fit to
experimental dispersion data (left). Shear wave velocity Vs profile from inversion (right).

Experimental 5.02 Hz
Theoretical 5.03 Hz

Figure 7.2.18. Manta Port parking area: Mean horizontal-to-vertical spectral ratio (HVSR) obtained.
The red lines are the theoretical H/V spectral ratio curve and peak for the developed Vs profile.

At the wharf deck, the Vs profile developed along with the experimental dispersion curve
with theoretical fit for the site are shown on Table 7.2.8 and Fig. 7.2.21. From the Vs profile, a
soft layer is encountered at the surface with a Vs of 290 m/s to a depth of 3 m below the ground
surface. Beneath that, a stiffer layer with a Vs of 487 m/s is encountered to a depth of 4.7 m,
followed by two softer layers, with Vs of 317 m/s and 228 m/s. The final layer of the profile
with a Vs of 320 m/s was encountered at a depth of 8.1 m.

Shear wave velocity data for three locations at the Manta Port from MAM and MASW tests
are presented in Fig. 7.2.22.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-20
Active MASW Array
Active Source Locations

Meters
0 123

Figure 7.2.19. Manta Port wharf deck before the earthquake with locations of active surface wave
arrays. Base image: Digital Globe (©2016).

Figure 7.2.20. Manta Port wharf deck post-earthquake: GEER team member C. Wood with an
Ecuadorian colleague at the active MASW array (GPS: 0° 56' 14.9" S, 80° 43' 28.6" W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-21
Table 7.2.8. Manta Port wharf deck: Median in-situ measured Vs profile.
Layer Depth to Bottom Layer Shear Wave
Number of Layer Thickness Velocity Vs
# (m) (m) (m/s)
1 3.0 3.0 290
2 4.7 1.8 487
3 5.2 0.5 317
4 8.1 2.9 228
3 25 16.9 320
Vs30 (m/s) NA Vs to depth of 30 m

Station: Station:
Manta Port 1 Latitude: -0.933 Latitude:
Manta Port 1 -0.933 Station: Station:
Manta Port 2 Latitude: -0.933 Latitude:
Manta Port 2 -0.933
City:
Figure
Manta City:
7.2.21. Manta
Longitude:
Manta
Port wharf deck: City:
-80.722Longitude: -80.722
MedianManta
Rayleigh
City:
wave dispersion
Longitude:
Manta
curve;-80.723
Station:
-80.723Longitude:
fit Port
Manta to 3experimental
Latitude: -0.935 Station

Vs30 (m/s): data (left).


495.90 Vs Median
(m/s): 495.90 of top 1,000 Vs profiles
30
from 301.10
Vs (m/s): > 500,000
Vs (m/s):models
301.10 in inversion search (right).
City:
Vs (m/s):
Manta Longitude: -80.723
30 30
City:
30 271.90 Vs30 (m

Shear wave velocity, Vs (m/s)


Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)
Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)
0 100 200 300 0 400100500200600300700400800500900600 700 800 900 0 50100 150 200 0 250 100
50 300 150
350 200
400 250
450 300
500 0350 50400100450150500200 250 300 350 400 450 800 10
0 0 0 0 0 0
Station: Manta Port 1Manta Latitude:
Station: Port 1 Latitude:
-0.933 -0.933 Station: Manta Port 2Manta Latitude:
Station: Port 2 Latitude:
-0.933 -0.933Station: Manta Port 3 Latitude: -0.935 St
5
City: City:
Manta Manta Longitude: Longitude:
-80.722 -80.722 City: City:
Manta Manta Longitude: Longitude:
-80.723 City:
-80.723 Manta Longitude: -80.723 Ci
5 5 5 5 5 10
Vs30 (m/s): Vs
495.90
30
30
(m/s): 495.90 Vs30 (m/s): Vs
301.10
30
30
(m/s): 301.10 Vs30 (m/s): 271.90 Vs
15
Station:
10 Port 1Manta Port Latitude:
1 10 Shear -0.933wave Vs (m/s) Station:
Manta Port 2Manta
10 Port 2 Latitude:
Shear -0.933wave
wave Shear Station:
velocity, (m/s) Manta
Vs velocity, Port 3 10 Latitude: Shear
Vs (m/s) -0.935wave velocity, Vs (m/s) Station: Catolica University La
Station: Manta Latitude: -0.933wave Shear
velocity, Vs velocity,
(m/s) Station: Latitude: 10
-0.933 20
City: City: Manta 0Longitude: Longitude: -80.722 City: City:
Manta Manta Longitude: Longitude:
0-80.723 -80.723
City: Manta 400 450 0Longitude: 150 200 250 300 350 City:
500 50 100-80.723 400 450 Guayaquil Lo
Manta -80.722
0 100 200 300 400 500
100 200 300 400 500 600 700 800 900 600 700 800 900 0 50 100 15050 200
100250
150300
200350250400
300 350500
450 800
0 0 0 0 0 25 0
Vs30 (m/s): Vs 30 (m/s): 495.90
495.90
15 15 Vs30 (m/s): Vs 30 (m/s): 301.10
301.10 15 15 Vs30 (m/s): 271.90 15 Vs30 (m/s): 1258.60
30 5
Shear waveShear wave
5
velocity, Vsvelocity,
(m/s) 5 Vs (m/s) Shear waveShear wave
5
velocity, Vsvelocity,
(m/s) 5 Vs (m/s) Shear wave velocity, Vs (m/s) Shear wave velocity,
35 V
5 10
0 20 400 20 50 100 20 20 400 450 0500 50
Depth (m)

200 20250

Depth (m)
100 400
200 500
300 600 500 800600 900
700 800 900 0 50 0
100 150 200 250 300 350 100 150 300 350 400 450 800 1000 1200 1400 1600 180
0 100 0200 300 700
0 0 150 200 250 300 350 400 450 500 0 0 40
10 10 10 10 15
45
Depth (m)

10
Depth (m)
Depth (m)

Depth (m)

25 25 25 25 25 5 20
5 5 5
15 15 15 15 5 50
10 25
55
10 30 30 10 10 30 30 30 15 15
20 20 20 20 30
60
10
20 35
15 15 65
Depth (m)

15 35 25 35
Depth (m)
Depth (m)

35 35 25
Depth (m)

Depth (m)

35 20
Depth (m)

25 25 25 40
70
15
20 20 20 30 45
40 30 40 30 40 30 30 40 40 25 75
35 50
80
Depth (m)
Depth (m)

25 25
Depth (m)

Depth (m)

20
Depth (m)

35 25 35 35
35 40
45 45 45 45 45 85 55
30
30 30 30 45 60
40 40 40 40 25
50 50 50 50 50
35 65
35 35 35
45 45 45 45 55
30 70
55 55 55 55
40 40 60 75
50 40 50 50 40
50
Figure 7.2.22. Manta Port: MAM and MASW Vs at Sites35 (Stations) 1, 2, 3 (GPS: 0° 55' 58.8" S, 80°
65 80
45 45 45
55 43' 19.2"
55
W, 0° 55' 58.8" S, 80° 43' 22.8"55 W, 0°5556' 6" S, 80° 43' 22.8" W). 45 70 85

50 50 40 75
50
80
55
GEER-ATC Muisne, Ecuador, Earthquake
55
7-Geotechnical Observations
55 45 85
Report Version 1 7-22
Mejia Bridge, Post-Earthquake Geotechnical Information

One boring was drilled post-earthquake at the Mejia Bridge (between site to a depth of 19
m from the surface (0° 59' 24.0" S, 80° 28' 12.0" W). Details on the bridge, located on the road
between Portoviejo and Crucita, were provided in the pre-earthquake part of this section.

The boring encountered ground water at a depth of 3.35 m. The subsurface profile, based
on the log of Fig. B3 in Appendix B, consist of a top 1.5-m mixed layer of silty and clayey
sands with energy-corrected N60 values from 11 to 27 blows/30cm, water content w from 22
to 32% and a FC between 20 and 43%. This layer is followed by 3.9 m of gravel with a N60
values from 10 to 31 bl/30cm, w between 7 and 15% and FC from 8 to 13%. Beneath that, to
a depth of 12.55 m, a silty sand with N60 varying from 5 to 13 bl/30cm, w between 16 and 34%
and FC from 12 and 61% was present. From depth between 12.55-14.05 m there is a sandy
clay layer with N60 between 45 and 48 bl/30cm, w of 22-29%, PI between 15 and 19, and FC
from 40 to 57%. The bottom part of the borehole to a depth of 19 m revealed silty sand with
N60 between 40 and 57 bl/30cm, w of 16-29% and FC between 14 and 23%. The adjacent CPT
(0° 59' 27.6" S, 80° 28' 12.0" W) was performed up to the same depth of 19 m, with results
presented on Fig. B4 of Appendix B.

Geophysical
Station: data from MAM
Mejia Bridge Latitude: and MASW tests are Station:
-0.991 presented
Mejiafor two sites
Bridge (Fig.
Latitude: 7.2.23),
-0.990
City: Portoviejo Longitude: -80.469 City: Portoviejo Longitude: -80.470
indicatingVsV(m/s):
s variability
176.50 to a 15-m depth, although they are
30 in close
Vs (m/s): 214.10proximity (Fig. 7.2.6). 30

Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)


0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350
0 0

10

10
15
Depth (m)
Depth (m)

20
15

25
Station: Mejia
Station: Mejia Bridge
Bridge Latitude: -0.991
Latitude: -0.991 Station: Mejia
Station: Mejia Bridge
Bridge Latitude: -0.990
Latitude: -0.990
20 City:
City: Portoviejo
Portoviejo Longitude:
Longitude: -80.469
-80.469 City:
City: Portoviejo
Portoviejo Longitude:
Longitude: -80.470
-80.470
VsVs 30 (m/s):176.50
30 (m/s):
176.50 VsVs 30 (m/s):
30 (m/s):
214.10
214.10
30

Station:
Station: Mejia
MejiaBridge
Bridge Latitude: Shear
Latitude:Shear
-0.991 wave
-0.991
wave velocity,
velocity, VsVs (m/s)Station:
(m/s) Station: Mejia
MejiaBridge
Bridge Latitude: Shear
Latitude:Shear
-0.990 wave
-0.990
wave velocity,
velocity, VsVs (m/s)
(m/s)
City:
City: Portoviejo
Portoviejo 0Longitude:
Longitude: 150150 200200 250250 300300 City:
-80.469
0 50 50 100 -80.469
100 City:
350350 400400 Portoviejo
Portoviejo 0Longitude:
Longitude:
0 50 50 -80.470
-80.470
100 150150 200200 250250 300300 350350
100
0 0 0 0
VsVs (m/s):
30 30 (m/s):176.50
176.50 25 VsVs (m/s):
30 30 (m/s):214.10
214.10 35

Figure 7.2.23. Mejia Bridge MAM and MASWShear


Shear
Shearwave
wavevelocity, VsVs
velocity, (m/s)
(m/s)
Vs wave
Shear velocity,
atwave
Sites (GPS:
1,Vs2Vs(m/s)
velocity, (m/s)
0° 59' 27.6" S, 80° 28' 8.4" W,
5 250
0 0
0 0
5050 100
0° 59' 24.0" S, 80° 28' 12.0" W).
100 150
150 200
200 250
250 300
300 350
350 4000 0
400
0 0
5050 100
100 150 5200
150 200 250 300300 350
350

5 5

GEER-ATC Muisne, Ecuador, Earthquake 10 10 7-Geotechnical Observations


5 5
Report Version 1 7-23
5 5
10 10
15 15
)
)
Las Chacras (Rio Chico) Bridge, Post-Earthquake Geotechnical Information

Surface wave measurements were taken at a relatively undamaged section of the SE


embankment of the Las Chacras (Rio Chico) Bridge (Figs. 7.2.24, 7.2.25). The Vs profile with
the experimental dispersion curve and theoretical fit are shown on Table 7.2.9 and Fig. 7.2.26.
Active MASW Array
Active Source Locations

Meters
0 66

Figure 7.2.24. Las Chacras (Rio Chico) Bridge before the earthquake with locations of the active
surface wave array. Base image: CNES/Astrium (©2016).

Figure 7.2.25. Las Chacras (Rio Chico) Bridge post-earthquake: GEER team member A. Zekkos with
an Ecuadorian colleague at the active MASW array (GPS: 0° 58' 34.4" S, 80° 25' 20.0" W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-24
At the top 4.5 m, a stiff layer with Vs of 380 m/s was detected, at the same height of the
embankment and within the range of observed roadway cracks, followed by a 4.5-m soft soil
layer with Vs of about 100 m/s, which could have failed, causing the embankment failure. A
somewhat stiffer layer with Vs of 200 m/s was found below the soft layer to a 12-m depth. The
bottom layer tested to a depth of 25 m below the embankment has Vs of 350 m/s.

Table 7.2.9. Las Chacras (Rio Chico) Bridge: Median in-situ measured Vs profile.
Layer Depth to Bottom Layer Shear Wave
Number of Layer Thickness Velocity Vs
# (m) (m) (m/s)
1 4.5 4.5 380
2 9.0 4.5 100
3 12.0 3.0 200
4 25.0 13.0 350
Vs30 (m/s) NA Vs to depth of 30 m
500 0
Mean with +/- 1 
450 Theoretical Fit
5
400

350
Phase Velocity (m/s)

10
300
Depth (m)

250 15

200
20
150

100
25
50

0 30
10 100 0 100 200 300 400
Frequency (Hz) Shear wave Velocity (m/s)

Figure 7.2.26. Las Chacras (Rio Chico) Bridge: Median theoretical Rayleigh wave dispersion curve;
fit to experimental dispersion data (left). Vs profile obtained from during inversion (right).

Chone Dam, Post -Earthquake Geotechnical Information

The Chone Dam (Fig. 7.2.1) is located in central Ecuador (0° 41' 41.7" S, 79° 59' 21.9" S).
The site experienced minor damage during the main earthquake without any operation
disruption. Surface wave measurements were conducted along: (i) the top of the dam, and (ii)
the mid-portion of the dam (Figs. 7.2.27 to 7.2.29). The experimental dispersion curve and
theoretical fit and associated Vs profile for both locations are shown in Fig. 7.2.310. The Vs
profiles are tabulated in Table 7.2.10 and 7.2.11.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-25
Figure 7.2.27. Chone Dam post-earthquake photo showing the locations of active and passive surface
wave arrays, and HVSR (GPS: 0° 41' 41.7" S, 79° 59' 21.9" S).

Figure 7.2.28. Chone Dam: Active MASW array at top of dam (GPS: 0° 42' 1.6" S, 79° 59' 22.0" W)
with reconnaissance team members A. Caicedo, K. Rollins, C. Wood and other local colleagues.

At the top of the dam, the testing identified a relatively soft layer with Vs of 239 m/s that
increases with depth due to confining stress, reaching the max Vs of 364 m/s at 38 m depth. At
the middle of the dam, the Vs profile indicates a top thin soft layer with Vs of 185 m/s. However,
GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations
Report Version 1 7-26
the stiffness picks up at 0.8 depth to Vs of 367 m/s, which is similar to the Vs measured from
the top of the dam at that depth. Although the mid-dam measurements were taken far from the
center of the dam, it is likely that the confining stress still affects the soil stiffness. Additionally,
the dam was designed for the soil profile under the middle to be slightly different than the one
at the center. At the middle, the profile reaches a max 50 m below the top of the dam that does
not extend to the foundation depth, primarily due to lack of passive energy in the vicinity. The
HVSR results are shown in Fig. 7.2.31 with a HVSR peak at around 3.4 Hz (~0.3 s) for the
middle of the dam profile indicating presence of a stiff layer below the bottom of the Vs profile.

Figure 7.2.29. Chone Dam Middle: Active MASW array (GPS: 0° 41' 56.4" S, 79° 59' 23.1" W).
Table 7.2.10. Chone Dam: Median in-situ measured Vs profile.
Layer Depth to Bottom Layer Shear Wave
Number of Layer Thickness Velocity Vs
# (m) (m) (m/s)
1 5.3 5.3 239
2 18.9 13.6 295
3 37.8 18.9 364
Vs30 (m/s) 304 Vs to depth of 30 m

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-27
Table 7.2.11. Middle of Chone Dam: Median in-situ measured Vs profile.
Layer Depth to Bottom Depth to Bottom of Layer Shear Wave
Number of Layer Layer from top of dam Thickness Velocity Vs
# (m) (m) (m) (m/s)
1 0.8 20.8 0.8 363
2 3.8 23.8 3.0 367
3 17.6 37.6 13.9 371
4 30.0 50.0 12.4 494
Vs30 (m/s) 412 Vs to depth of 30 m

Figure 7.2.30. Chone Dam: Median theoretical Rayleigh wave dispersion curve; fit to experimental
dispersion data (left). Median of top 1,000 Vs profiles from > 500,000 during inversion (right).

4
Experimental HV MOD
Experimental HV TOD
3
H/V Spectral Ratio

0 0 1
10 10
Frequency (Hz)
Figure 7.2.31. Chone Dam: Mean horizontal-to-vertical spectral ratio (HVSR). Theoretical lines
representing the H/V ratio curve and peak for median of top 1,000 profiles (next version of this report).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-28
APO1 Strong Motion Station, Post -Earthquake Geotechnical Information

Boring P-1 (1° 2' 13.2" S, 80° 27' 32.4" W) was drilled at the accelerometer station APO1
to a depth of 27.1 m, where siltstone rock was encountered (log in Fig. B5 of App. B). The top
7.25 m of the profile is of a mix of clays and silts with N60 varying from 5 and 40 bl/30cm, w
from 15 to 53%, PI between 20 and 38, and FC between 67 and 94%. This layer is followed
by 1 m of sand, and another mix of clays and silts to 22.6 m, with N60 from 19 to 100 bl/30cm,
w between 29 and 71%, PI variation from 37 to 54, and FC from 81 to 96%. Dark brown
siltstone was encountered from to the final depth, cored using rotation drilling since ground
water was not present. The siltstone cores (Fig. 7.2.32) were subjected to laboratory Uniaxial
Compression tests that resulted in an average qu of 12 MPa. Geophysical MAM and MASW
results on Fig. 7.2.33 show Vs > 500 m/s below 22.5 m, consistent with siltstone presence.

Station: APED Latitude: 0.068 Station: APO1 Latitude: -1.038


City: Pedernales Longitude: -80.057 City: Portoviejo Longitude: -80.460
Vs30 (m/s): 348.30 Figure 7.2.32. Siltstone cores from
Vs30 (m/s): boring P-1 at the APO1 accelerometer station.
240.00

Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)


0 100 200 300 400 500 600 700 800 0 100 200 300 400 500 600 700 800 900
0 0
Station: APED Latitude: 0.068 Station: APO1 Latitude: -1.038
5 5
City: Pedernales Longitude: -80.057 City: Portoviejo Longitude: -80.460
0 Vs30 (m/s): 348.30 10 Vs30 (m/s): 240.00
5 15
Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)
0 0 100 200 300 400 500 600 20700 800 0 100 200 300 400 500 600 700 800 900
0 0
5 25
5 5
0 30
10 10
5 35
15 15
Depth (m)

0 40
20 20
5 45
25 25
0 30 50 30

5 35 55 35
Depth (m)
Depth (m)

0 40 60 40

5 45 65 45

0 50 70 50

5 55 75 55
Latitude: 0.068 Station: APO1 Latitude: -1.038
60 80 60
0 Longitude: -80.057 City: Portoviejo Longitude: -80.460
es
65 Vs30 (m/s): 240.00 65
5 85

ve velocity, Vs (m/s)
70
Figure 7.2.33. APO1
Shear station MAM
wave velocity, and MASW
Vs (m/s)
70
Vs (GPS: 1° 2' 16.8" S, 80° 27' 36.0" W).
75 75
400 500 600 700 800 0 100 200 300 400 500 600 700 800 900
0
GEER-ATC
80 Muisne, Ecuador, Earthquake 80 7-Geotechnical Observations
Report
85 Version 1
5
85 7-29
10

15
AMNT, ACHN, APED, AMA1 Stations, Post-Earthquake Geotechnical Information

Geophysical Vs data from MAM and MASW at the accelerometer stations AMNT (Manta),
Station: AMNT APED Latitude: Station:
-0.941-0.941 and AMA1Station: ACHN ACHN Latitude:
Latitude: -0.698
ACHN (near
Station: Chone),
AMNT (Pedernales),
Latitude: (Esmeraldas) are shown on Fig.
-0.6987.2.34
City: City: MantaManta Longitude:
Longitude: -80.735
-80.735 City: City: Chone Chone Longitude:
Longitude: -80.084-80.084
with ground motions recordings provided in Chapter 5.
Vs30 (m/s):
Vs30 (m/s): 513.00513.00 Vs30 (m/s):
Vs30 (m/s): 208.90208.90

ShearShear
wave wave velocity,
velocity, Vs (m/s)
Vs (m/s) Shear Shear wave velocity,
wave velocity, Vs (m/s)
Vs (m/s)
0 0 200 200400 400600 600800 800
1000 1000
1200 1200 0 0
50 50
100 100
150 150
200 200
250 250
300 300
350 350
400 400
0 0 0 0

5 5 5 5

10 10 10 10

15 15 15 15

20 20 20 20

25 25 25 25

30 30 30 30

35 35 35 35

Depth (m)
Depth (m)

Depth (m)
Depth (m)

40 40 40 40

45 45 45 45

50 50 50 50

55 55 55 55

60 60 60 60

65 65 Station:
Station: AMNT
AMNT Latitude:
Latitude: -0.941
-0.941 65 Station:
Station: ACHN
65 ACHN Latitude:
Latitude: -0.698
-0.698
City:
City: Manta
Manta Longitude:
Longitude:-80.735
-80.735 City:
City: Chone
Chone Longitude:
Longitude:-80.084
-80.084
70 70 70 70
Vs
Vs3030(m/s):
(m/s): 513.00
513.00 VsVs (m/s):
30 30 (m/s):208.90
208.90
75 75 75 75
Station: Latitude: Station:
Station:
Station: AMNT
AMNT Station:
Latitude: APED
-0.941
Latitude: Shear wave
-0.941
Shear Latitude:
wavevelocity, Vs
velocity, Vs(m/s) 0.068
(m/s) Station:
Station: ACHN
ACHN Station:
Latitude:
Latitude: AMA1 APO1
-0.698
Shear wave
-0.698
Shear velocity,
wave VsVs
velocity, (m/s) 0.943 -1.038
Latitude:
(m/s)
80 80 City: 80 80City: Esmeraldas
City:
City: Manta 0Longitude:
Longitude: -80.735
Pedernales Longitude: City:
-80.057
City: Chone City:
0Longitude:
Longitude: -80.084
Portoviejo
-80.084 Longitude: -79.731
Longitude: -80.460 City:
Manta 0 200
200 -80.735
400
400 600
600 800
800 1000
1000 1200 Chone
1200
0
50 100
0 150
50 200 250
100 300
150 350
200 400
250 300 350 400
00 0 (m/s):
Vs
Vs30 (m/s): 513.00 85 85 Vs (m/s): 348.30 Vs
Vs30 (m/s):
(m/s): 208.90 85 Vs30 333.00240.00
85Vs (m/s): Vs30 (m/s)
30 (m/s): 513.00 30 208.90 30 30
55 5 5
Shear
Shearwave
wavevelocity,
velocity,Vs
Vs(m/s)
Shear wave velocity, Vs (m/s)
(m/s) Shear
Shearwave
wavevelocity, VsVs
velocity, (m/s)
ShearShear
(m/s) wave wave
velocity, Vs (m/s)
velocity, Vs (m/s)
10
10 1010
0 50 100 150 200
0
00 0
200
200 400
400 600
600 0 800
800
100
1000 1200
1000 300
200 600 0 0
1200400
500 700 50 800 100 150 200 0 250 250
100 300
0 200300 350
100 300 200
400
350 400 400500400
300 600500 700 600
800 700
900 1000
800 900 0 100
015 0 0 1515 0 0
15
55 55
520
20 5 2020 5
5
10 1010
10 25
10 25 10 252510
15 15 30 10
15 30 15
15 30 15 3015
20 20 35
20 35 20
20 35 20 3520 15
Depth (m)
(m)

Depth (m)

25
Depth(m)

25 40 40
25 25
25 40 25 4025 20
Depth

30 45 30 45
30 45 30 45
30 30 30
35 35 50
35 50 35 25
50 50
Depth (m)
(m)

35 35 35
Depth (m)
Depth(m)

40 40 55
40 55 40
Depth (m)
Depth (m)
Depth (m)
Depth (m)

55 55 30
Depth

45 40 45 40 40
45 60 45 60
60 60
50 45 50 45 65 45 35
50 65 50
65 65
55 5070 55 50 70 50
55 55 70 40
70
60 5575 60 55 75 55
60 60 75
75
65 65 45
65 6080 65 60 80 60
80 80
70 Station: APED Latitude: 70 0.068 Station:
Station: APO1
AMA1 Latitude:
Latitude: -1.038
0.943 Station: AV1
70 6585 70 65 85 65 50
85City: Pedernales Longitude: -80.057 85
City:
City: Portoviejo
Esmeraldas Longitude:
Longitude: -80.460
-79.731 City: Quin
75 75
75 70 Vs (m/s): 348.30 75 70 Vs 70 (m/s):
Vs Vs30 (m/s): 283
30 3030 (m/s): 240.00
333.00
55
80 80
80 75 80 75 75
85 Station: APED Latitude: Shear
0.068wave velocity, Vs (m/s)85Station:
Station: APO1
AMA1 Latitude: Shear
Latitude: Shear
-1.038
0.943
wave
wavevelocity,
velocity,VsVs(m/s)
(m/s) Station: AV18 Latitude: Shea
0.31
60
85 80 0Longitude:
100 200 300 400 500 600 85
700 800 80 0 080 100
100 200 300
300 400
400 500
500 600
600 700700 800
900 900
800City: 1000 Quininde 0 100 200
City: Pedernales -80.057 City:
City: Portoviejo
Esmeraldas Longitude:
Longitude: -80.460
-79.731 0
Longitude: -79.
0 00
Vs30 (m/s): 348.30 85 Vs
Vs3030(m/s):
(m/s): 240.00
333.00 85 85 Vs30 (m/s): 283.00 65
5 55
5
Shear
Shearwave
wavevelocity,
10 VsVs(m/s)
Shear wave velocity,
10 Vs (m/s)
Figure 7.2.34. MAM/MASW at stations AMNT, ACHN, velocity,
APED,
10 AMA1 (clockwise from top Shear
(m/s) wave velocity, Vs (m/s)
left:
0 100 200 300 400 500 600 700 800 00 100
100 200 300
300 400
400 500
500600600700700
800 800
900 900
1000 0 100 200 300 400 10500 600 700
0 0°56'27.6"S,
15 80°44'6"W; 0°41'52.8"S,0080°5'2.4"W; 0°4'4.8"S, 80°3'25.2"W; 0°56'6"S, 79°43'20"W).
15
15 0

5 20 55 20
20 15
5
10 GEER-ATC 25
Muisne, Ecuador, Earthquake 10
10 25
25 7-Geotechnical Observations
20
Report Version
30
1 30
30
10 7-30
15 15
15
25
20 35 20
20 35
35 15
)
)
)
)
AV18, ALIB, AMIL, AGYE Stations, Post-Earthquake Geotechnical Information

de: 0.943 Geophysical


Station: Vs data fromLatitude:
AV18 MAM and 0.312 MASW at the stations
Station: AV18 (Quininde),
ALIB Latitude: ALIB (La
-2.243 Station:
tude: -79.731 City: Quininde Longitude: -79.478 City: Sta. Elena Longitude: -80.849 City:
Libertad), AMIL (Milagro), and AGYE (Guayaquil) are
Vs (m/s): 283.00
shown on Fig. 7.2.35.
Vs (m/s): 433.00 Vs30 (m/s)
30 30

m/s) Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)


0 800 900 1000 0 100 200 300 400 500 600 700 800 0 100 200 300 400 500 600 700 800 900 1000 0 100
0 0 0

5 5
5
10 10
10
15 15
15 20 20

20 25 25

30 30
25
35 35
Depth (m)

Depth (m)
Depth (m)
30 40
40

35 45 45

50 50
40
55 55
45
60 60
Latitude: 0.943 Station: AV18 Latitude: 0.312 Station: ALIB Latitude: -2.243 Station: A
50 65 65
Longitude: -79.731 City: Quininde Longitude: -79.478 City: Sta. Elena Longitude: -80.849 City: M
Vs30 (m/s): 283.00 70 Vs30 (m/s): 433.00 70 (m/s): 37
Vs
55 30

Latitude: Shear 75 75
ty, Vs (m/s) Station: AV18 0.312wave velocity, Vs (m/s) Station: ALIB Latitude: Shear wave velocity, Vs (m/s)
-2.243 Station: AMIL Latitude: She
-2
de: 60 Station: Latitude: -2.181
-2.243
600 700 800 City:
900 1000 Quininde 0 100 AMIL
Longitude: -79.478
200 300 400 500 600 700City:800 Sta. Elena 80 0Longitude: 300 400 500 600 700 800 City:
100 200 -80.849 900 1000 Milagro 080
Longitude:
100 200-7
0 0 0
ude: -80.849Vs30 (m/s): 283.00 City: Milagro Longitude: -79.529
Vs30 (m/s): 433.00 Vs30 (m/s): 370.00
65 85 85
5 5
30 Vs (m/s) Vs5 (m/s): 370.00
Shear wave velocity, Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)
10 10
m/s) 0 100 200 300 400 10 500 600Shear
700 800 velocity, Vs (m/s)
wave 0 100 200 300 400 500 600 700 800 900 1000 0 100 200 300 400 500 600 700 8
0 0 15 0 15
800 900 1000 0 100 200 300 400 500 600 700 800 900
5 0 15 5 20 5 20

5 20 10 25 10 25
10
15 30 15 30
10 25
15 20 20
35 35
15
Depth (m)

Depth (m)
Depth (m)

30 25 25 40
20 40
20
35 30 45 30 45
25
25
35 50 35 50
Depth (m)

40
Depth (m)
Depth (m)

30 30 40 55 40 55

35 35 45 45 60 45 60
Depth (m)

40 50 50 65 50 65
40
45 55 55 70 55 70
45
60 75 60 75
50
60
50 65 80 65 80
55
65 70 85 70 85
55
60
75 75
Latitude:60 -2.243 65 Station: AMIL Latitude: -2.181
80 80
Longitude: -80.849 City: Milagro Longitude: -79.529
65 70 85 85
Vs30 (m/s): 370.00
75
ty, Vs (m/s) Station: AMIL Latitude: -2.181wave velocity, Vs (m/s)
Shear
City: 80 Longitude: -79.529
600 700 800 900 1000 Milagro 0 100 200 300 400 500 600 700 800 900
0
Vs30 (m/s): 370.00 85
5
Figure 7.2.35. MAM/MASW at AV18, ALIB, AMIL, AGYE (clockwise from top left: 0°18'46.8"S,
Shear wave velocity, Vs (m/s)
10
0
0 100 200 79°28'40.8"W;
300 400 500
15
600 2°14'34.8"S,
700 800 900 80°50'45.6"W; 2°10'51.6"S, 79°31'44.4"W; 2°3'14.4"S, 79°57'7.2"W).
5 20

10 GEER-ATC Muisne,
25
Ecuador, Earthquake 7-Geotechnical Observations
Report Version 1 7-31
15 30

20 35
AGYE1, AGYE2, Catolica University Stations, Post-Earthquake Geotechnical Information
Station: Manta Port 3 Latitude: -0.935 Station: Catolica University Latitude: -2.181
City: MAMManta
and MASW Vs data-80.723
Longitude: from Guayaquil stations
City: AGYE1
Guayaquil
and Catolica
Longitude: University
-79.904
are
Vs (m/s): 271.90 Vs30 (m/s): 1258.60
shown
30
on Fig. 7.2.36.
Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)
0 50 100 150 200 250 300 350 400 450 800 1000 1200 1400 1600 1800 2000 2200 2400
0 0

5
5 10

15
10
20

25
15
30

35
Depth (m)

20

Depth (m)
40

45
25
50

55
30
60

35 Station: Latitude: 65
Manta Port 3 -0.935 Station: Catolica University Latitude: -2.181
City: Manta Longitude: -80.723 70 City: Guayaquil Longitude: -79.904
Vs30 (m/s): 271.90 Vs30 (m/s): 1258.60
40 75
Station: Manta Port 3 Latitude: -0.935 Station: Catolica University Latitude: -2.181
Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)
City: Longitude: -80.723 80 Longitude: -79.904
Manta City: Guayaquil
0 50 100 150 200 250 300 350 400 450 800 1000 1200 1400 1600 1800 2000 2200 2400
Vs30 (m/s): 271.90
45 0 Vs30 (m/s): 1258.60 85 0

5
Shear wave velocity, Vs (m/s) Shear wave velocity, Vs (m/s)
0 50 100
Figure 5 7.2.36. AGYE1 and Catolica
150 200 250 300 350 400 450
University stations MAM and MASW Vs measurements
800 1000 1200 1400 1600 101800 2000 2200 2400
0 (2°15'0.0"S, 79°54'36.0"W; 2°10'51.6"S,
0 79°54'14.4"W). 15
10 5
20
5
Strong motion station AGYE2
10 is at the Ramon Unamuno Stadium. It is a typical estuarine
25
15 15
10 deltaic Guayaquil deep soil profile
20
with Vs30 of about
30
100 m/s and groundwater at 2.5 m depth.
35
The V20s values shown on Fig. 7.2.37 were measured in-situ or correlated with conventional
Depth (m)

Depth (m)

25
40
15
30
45
geotechnical
25 testing results from
35 previous investigations in 2005 and 2014 (Vera-Grunauer,
50
Depth (m)

20
Depth (m)

40
2014).30The 2005 in-situ tests included SPT, sCPT, Vane
55 Shear, VST, and Spectral Analysis of
45
25 60
Surface Waves, SASW. In 2014,
50 a deep geotechnical boring, ERU, reaching almost 145 m
35 65
55
30 below the ground surface was drilled with SPT 70terminated at 37.5 m where refusal was
60
40 75
35 encountered (N60 > 60 bl/30cm)65 and drilling followed
80
with rotary methods to 107 m. CPTu
70
was performed
45 to 39 m. Geophysical
75
Multi-channel85Analysis of Surface Wave with Refraction
40

Microtremor tests (MASW and 80ReMi) were also performed.


45 85

The AGYE2 profile has been analyzed in the next section of this report on site response. It
consists of 1 m of granular fill which overlays a mixed clay and sand stratum that is divided in
2.5 m of dark yellowish clay, followed by 16.5 m of clay with thin seams of silty sand and 16
m greenish gray stiff clay, ending with 4 m of hard clay. The PI of this deposit varies from 40

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-32
to 60 up to 5 m and from 20 to 40 up to 38 m. Below 40 m, there is very dense 12-m sand layer,
followed by a greenish gray hard clay at depth from 52 to 80 m with unconfined compression
strength, qu ≥ 500 kPa. About 25 m of very dense sand underlays the clay, followed by greenish
gray hard clay with sporadic presence of gravel.
Vs (m/s)

PRES. HID.
(t / pulg )
2
COLOR
UNIFIED SOIL
0 200 400 600 800 1000 1200 1400 CLASSIFICATION SYSTEM

0
Estadio Ramón Unamuno, ERU 2014
CH
VsA = 120 (s'o/Patm)0.273 (Soft clay) ; Lin et al., 2014
Gmáx = CG3 Cub1 ex (so'/Pa)nG (Granular geomaterials) ; Menq, 2003 MH
CH
Vs = 280 (s'o/Patm)0.261 (Dense sand) ; Lin et al., 2014 SM
CH
SC
VsB = 230 (s'o/Patm)0.261 (Hard clay) ; Lin et al., 2014 CL
SM
CH
SM
Gmáx = Pat 380 e-1.3 (P/Pat)0.50 (Calibrated for GYE-clay) ; CH
modified of Pestana and Salvati, 2006
MH
Profile Selected
210 ECU (SASW at Puerto Azul, 2005)
20 216 ECU (SASW at Estadio Ramón Unamuno, 2005)

Young estuarine deltaic CH

Fine-grained soils
CL

SM

40
Poorly-graded SAND with SILT;
very dense;
dark gray

Dense Sand Poorly-graded SAND;


very dense;
dark gray

Poorly-graded SAND with SILT;


1.00 very dense; dark gray

SANDY CLAY; hard;


greenish gray

60
Hard Clay (qu>0.5MPa)
CLAY; hard;
greenish gray

SILT; grreenish gray


Depth (m)

CLAY; hard;
greenish gray

80
Well graded SAND with SILT;
very dense; dark gray

Very Dense Sand 1.50


Poorly graded SAND with SILT;
very dense; dark gray
Well graded SAND with SILT;
very dense; dark gray

100
SAND with SILT;
very dense; dark gray

GRAVEL with SAND and SILT;


4 cm; subrounded

CLAY: hard; greenish gray

Hard Clay (qu>0.5MPa) SILT; hard; greenish gray

120
CLAY: hard; greenish gray

GRAVEL with SAND and SILT;


4 cm; subrounded

Poorly graded SAND with SILT;


very dense; dark gray
GRAVEL with SAND and SILT;
4 cm; subrounded

Soft Claystone
SAND with SILT; dark gray
1.75

Soft CLAYSTONE

140

HALFSPACE, Vs=2800 m/s

Figure 7.2.37. AGYE2 station (GPS: 2°11'56.4"S, 79°53'52.8"W). s and soil profile based on previous
subsurface investigations in 2005 and 2014 (Vera-Grunauer, 2014).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-33
AMNT Strong Motion Station, Post-Earthquake Geotechnical Information

The Manta AMNT station is located in the northern part of the city, along Calle M1 road
by the coast. The location of the surface wave arrays and HVSR sensor are shown in Fig. 7.2.38
with the active MASW array on Fig. 7.2.39.
Active MASW Array
Passive 2D MAM Array
Active Source Locations
Horz to Vert Spectral Ratio
Strong Motion Station

Meters
0 44

Figure 7.2.38. Manta AMNT station before the earthquake with locations of active and passive surface
wave arrays, HVSR, and AMNT. Base image: Digital Globe (©2016).

Figure 7.2.39. Manta AMNT station active MASW array (GPS: 0° 56' 27.6"S, 80° 44' 4.9"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-34
The soil profile consists of a top 1.7-m thick layer with Vs of 193 m/s that varies at different
points around the site, which may indicate that its thickness varies spatially. Below to a depth
of 3.5 m is a stiffer soil layer with Vs of 391 m/s, followed by a stiff soil or highly weathered
rock layer with Vs of 480 m/s to 31.6 m, where bedrock of Vs  820 m/s was found. The Vs30
is 437 m/s and the fundamental site frequency of the profile above bedrock is 3.5 Hz (0.29 s),
based on Vs / 4H. The results of the HVSR testing are shown in Fig. 7.2.39 and the experimental
dispersion curve and theoretical fit with Vs profile on Figs. 7.2.40. A slight HVSR peak was
identified at approximately 12 Hz (0.08s). However, this peak did not fit the experimental
dispersion data, likely due to a thicker soil deposit under the HVSR location compared to the
surface wave array locations. Median Vs values are summarized on Table 7.2.12.

Figure 7.2.40. Manta AMNT station: Median Rayleigh wave dispersion curve; fit to experimental
dispersion data (left). Median of top 1,000 Vs profiles from > 500,000 models in inversion search (right).
11.96 Hz
2.71 Hz
11.96 Hz
2.71 Hz

Figure 7.2.41. Manta AMNT station: Mean horizontal-to-vertical spectral ratio (HVSR) obtained. The
red lines are the theoretical H/V spectral ratio curve and peak for the median of the top 1,000 profiles.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-35
Table 7.2.12. Manta SMS: Median in-situ measured Vs profile.
Layer Depth to Bottom Layer Shear Wave
Number of Layer Thickness Velocity Vs
# (m) (m) (m/s)
1 1.7 1.7 193
2 3.5 1.8 391
3 31.6 28.1 480
4 Bedrock Bedrock 820
Vs30 (m/s) 437 Vs to depth of 30 m
VsBR (m/s) 439 Vs to depth of bedrock
fsite (Hz) 3.47 Fundamental site frequency

Manta HV Measurements, Post-Earthquake Geotechnical Information

Horizontal to vertical spectral ratio (HVSR) measurements were taken at 7 locations (HV1-
HV7) of Fig. 7.2.42 at the central part of Manta, where heavy damage was observed following
the main event. HVSR plots are shown on Fig. 7.2.43 with peaks on Table 7.2.13. Except for
HV2 and HV3, all sites exhibited a low peak between frequencies 1.2-2.4 Hz (0.83-0.42 s).
Plotting these values on the site map in Fig. 7.2.44, the peaks indicate a basin that is shallower
near HV4 and HV1 and deeper at HV6. The absence of peaks at HV2 and HV3 may indicate
different subsurface conditions (changes in impedance between layers) under these locations.
Horz to Vert Spectral Ratio

Meters
0 384

Figure 7.2.42. Central Manta aero photo with locations of the seven HVSR test sites, HV1 to HV7.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-36
10 11.96 Hz
2.71 Hz HV1
HV2
8
HV3
H/V Spectral Ratio
HV4
6 HV5
HV6
4 HV7

0
1 10
Frequency (Hz)
Figure 7.2.43. Central Manta HVSR plots for sites HV1 to HV7.

Figure 7.2.44. Central Manta aero photo with locations of the HVSR tests and peaks for each site.
Table 7.2.13. Central Manta: Summary of measured HVSR peaks.
Site HVSR Peak (Hz)
HV1 1.8
HV2 NA
HV3 NA
HV4 1.8
HV5 2.4
HV6 1.2
HV7 2.7

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-37
IESS Hospital, Post-Earthquake Geotechnical Information

The IESS hospital is in southern Manta (0° 57' 17.8" S, 80° 43' 25.8" W) along Avenida
De La Cultura Road (Fig. 7.2.45). The hospital was heavily damaged in the main event and
lost its functionality. The array used by GEER for MASW testing is shown on Fig. 7.2.46.
Active MASW Array
Passive 2D MAM Array
Active Source Locations
Horz to Vert Spectral Ratio

Meters
0 85

Figure 7.2.45. IESS Hospital before the earthquake with locations of active and passive surface wave
arrays and HVSR. Base image: Digital Globe (©2016).

Figure 7.2.46. IESS Hospital active MASW array (GPS: 0° 57' 15.5" S, 80° 43' 25.8" W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-38
The dispersion curve with the developed Vs profile are shown on Figs. 7.2.47 and 7.2.48,
and Vs values on Table 7.2.14. The site has a soft top layer with a Vs of 178 m/s that extends
to 8.5 m below the surface, followed by a 288-m/s layer to 51 m depth where a weathered rock
layer was encountered with Vs of 1,020 m/s. This layering generated one major HVSR peak
and one minor HVSR peak, as shown on Fig. 7.2.48. The peaks at 1.44 and 5.0 Hz (0.7 and 0.2
s) match the theoretical peaks from the median developed Vs profile. Therefore, soil
amplification would be expected in these two frequency ranges. For the 5-story IESS hospital,
a rough estimate of its natural frequency is ~ 10 divided by the number of stories, or 2.0 Hz
(0.5 s), which is close to the HVSR peak of 1.44 Hz (0.7 s). Therefore, the ground motion was
likely amplified by the soil that was in resonance with the structure, contributing to the damage.

Figure 7.2.47. IESS Hospital: Median Rayleigh wave dispersion curve; fit to experimental dispersion
data (left). Median of top 1,000 Vs profiles from > 500,000 models in inversion search (right).
11.96 Hz
2.71 Hz

1.44 Hz 1.47 Hz

Figure 7.2.48. IESS Hospital: Mean horizontal-to-vertical spectral ratio (HVSR) obtained. The red
lines are the theoretical H/V spectral ratio curve and peak for the median of the top 1,000 profiles.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-39
Table 7.2.14. Median Vs profile at IESS Hospital.

Layer Depth to Bottom Layer Shear Wave


Number of Layer Thickness Velocity Vs
# (m) (m) (m/s)
1 8.4 8.4 176
2 51.0 42.5 287
3 Bedrock Bedrock 1020
Vs30 (m/s) 244 Vs to depth of 30 m
VsBR (m/s) 260 Vs to depth of bedrock
fsite (Hz) 1.28 Fundamental site frequency

Mobil, Post-Earthquake Geotechnical Information

Two boreholes, P-1 (0° 56' 52.8"S, 80° 43' 15.6"W) and P-2 (0° 56' 49.2"S, 80° 43' 8.4"W),
were drilled for reconnaissance at the Mobil site in Guayaquil. The boring logs can be found
in Figs. B6 and B7 of Appendix B.

Borehole P-1 reached a depth of 3.65 m with ground water table at 2.3 m. From the surface
up to a depth of 2.50 m the soil was a dark gray silty sand with gravels with a N60 variation
between 2 and 25 bl/30cm, w from 8 to 31% and FC from 17 to 29%. Between 2.5-3.15 m, P-
1 showed a yellow sandy silt with gravel and energy-corrected N60 between 3 and 5 bl/30cm,
w of 41-48%, and FC from 57 to 66%. In the bottom 0.5 m, a silty sand with gravel and N60
between 51 and 100 bl/30cm, w of 20-26% and FC between 7 and 31% was found.

The depth of Borehole P-2 was 4.7 m with ground water found at a depth of 2.0 m. The
boring log indicates a top 0.2 m of asphaltic concrete, followed by a mix of silty sands, silty
gravels and well-graded gravels with silt and sand to a depth of 4.25 m with N60 from 3 to 17
bl/30cm, w between 9 and 1% and FC from 7 to 18%. From the depth of 4.25 m to the bottom
of the borehole, a dark gray poorly graded gravel with silt and sand, and N60 of 100 bl/30cm,
w of 9% and FC of 6% was found.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-40
7.3 Site effects
INTRODUCTION

The main April 16th Mw7.8 earthquake was recorded by 30 national Strong Motion Stations
(SMS) with 90 records of the 3 components of the ground motion. As described in Chapter 5,
3 the organizations that own the SMS are: RENAC (Red Nacional de Acelerógrafos), OCP
(Oil Pipeline Network) and LMI (a collaboration of IG-EPN with the French IRD, Institute of
Research for Development). The maximum horizontal Peak Ground Acceleration (PGA)
recorded was 1.4 g at the APED SMS in Pedernales, as shown in Fig. 7.3.1, along with main
SMS locations, presented as color-coded circles according to the geo-mean PGA recorded in the main
event.

Figure 7.3.1. Strong Motion Stations (SMS) presented as color-coded circles according to the recorded
geo-mean PGA in the main event, with EW acceleration time histories and major cities of Ecuador.

As described in detail in the structural observations Chapter 8, after the main earthquake,
the Ministry of Housing evaluated about 33,667 buildings using ATC-20 procedures. Color-
coded results are presented by city on Fig. 7.3.2 (green: minor; yellow: partial; red: major).
Among the most affected cities were Portoviejo, Pedernales, Manta, Bahia de Caráquez and
Chone. Figure 7.3.3 shows average damage estimate as a percentage for each of these cities,
depending on the height or number of stories of the building stock, which is indicative of the
fundamental structural periods.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-41
Figure 7.3.2. Earthquake damage assessment of main cities after the main event using ATC-
20 by the Ministry of Housing. Colors indicate minor (green); partial (yellow); and major (red)
damage.

Figure 7.3.3. Average structural damage estimate expressed as a percentage among the 4 main
affected cities of Portoviejo, Pedernales, Manta, and Bahia de Caráquez, as function of the
number of floors. Data based on the Ministry of Housing post-earthquake assessment.

To gain insight in the effects of soil conditions to the ground motions and associated
damage, this section presents site effects based on actual records or on analysis by our team
based on reconnaissance data.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-42
AMNT STATION, MANTA

Dynamic Site Characterization Results

The Manta site of the AMNT SMS the GEER-ATC team performed post-earthquake
geophysical tests including MASW, MAM, and HVSR, to evaluate the in-situ Vs and
subsurface stratigraphy. Details on the field methods and data reduction for these tests are given
in the previous section with locations of the employed arrays on Fig. 7.3.4.

Figure 7.3.4. Manta AMNT station before the earthquake with locations of active and passive surface
wave arrays and HVSR (GPS: 0° 56' 27.6"S, 80° 44' 4.9"W). Base image: Digital Globe (©2016).

Figure 7.3.5 shows the raw experimental dispersion curves obtained from active MASW
and passive MAM (processed using HRFK and standard F-K). The dispersion curves are in
agreement in the 10-20 Hz range, indicating accuracy of the measurements. However, there is
a slight discrepancy between the passive and active dispersion curves at frequencies > 20 Hz.
This is likely due to variations in the thickness of the top layer across the site. For inversion,
the active dispersion curve was used for frequencies > 20 Hz due to higher near surface
resolution and closer proximity of the array to the SMS. For frequencies < 10 Hz, a slightly
higher phase velocity is observed using the HRFK than the standard F-K method. However,
the standard deviations overlap across the measured frequency range and therefore, the two
curves were averaged to develop the target dispersion curve for inversion, shown on Fig. 7.3.6.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-43
Figure 7.3.5. Manta AMNT station: Comparison of active-source MASW and passive-source 2D L-
array MAM dispersion data (HRFK and F-K processing), used as raw input for the Dinver inversion.

All dispersion curves are assumed to correspond to the fundamental mode of propagation.
For the inversion process, a system of 3 layers and halfspace was chosen, and no inverse
velocity layers were allowed that would create significant variations. Different layer
combinations were tried, but the solution converged on a simplified 4-layer system. However,
given the geological site conditions, a normally-dispersive profile was considered most likely.
After the inversion process which examined 500,000 models, the top 1,000 were selected and
their theoretical dispersion was fit to the experimental target curve (Fig. 7.3.6), with resulting
median Vs plotted on Fig. 7.3.7. The top 1,000 profiles had nearly identical soil layering and
the Vs indicated a limited solution space for the parameterization allowed.

Figure 7.3.6. Manta AMNT station: Theoretical Rayleigh wave dispersion curves for top 1,000 velocity
models from inversion. Theoretical dispersion curves with lowest misfit profile from the 1,000 profiles.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-44
The soil profile consists of a top soft layer with Vs of 193 m/s with thickness of 1.7 m that
likely varies spatially, causing discrepancies in the measured data. Below and up to 3.5 m depth
there is a stiffer layer with Vs of 391 m/s, followed by a stiff layer with Vs of 480 m/s to 31.6
m, where bedrock of Vs  820 m/s was encountered. The Vs30 is 437 m/s with fundamental site
frequency of 3.5 Hz (0.29 s), derived from Vs / 4H. HVSR results are shown on Fig. 7.3.8, with
a slight peak at approximately 12 Hz. However, this peak did not fit the experimental dispersion
data, possibly due to a thicker soil deposit under the HVSR location compared to the surface
wave array locations. Median Vs values are summarized on Table 7.3.1.

Figure 7.3.7. Manta AMNT station: Median Vs profile based on the top 1,000 profiles (lowest misfit
of ~0.413) from over 500,000 models during inversion.
Table 7.3.1 Manta AMNT station: Median in-situ measured Vs profile.
Depth to Bottom Layer Shear Wave
Layer
of Layer Thickness Velocity Vs
Number #
(m) (m) (m/s)
1 1.7 1.7 193
2 3.5 1.8 391
3 31.6 28.1 480
4 Bedrock Bedrock 820
Vs30 (m/s) 437 Vs to depth of 30 m
VsBR (m/s) 439 Vs to depth of bedrock
fsite (Hz) 3.47 Fundamental site frequency

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-45
11.96 Hz
2.71 Hz 11.96 Hz
2.71 Hz

Figure 7.3.8. Manta AMNT station: Mean horizontal-to-vertical spectral ratio (HVSR) obtained. The
red lines are the theoretical H/V spectral ratio curve and peak for the median of the top 1,000 profiles.

Ground Motion Analysis

The AMNT SMS is at the edge of a 24.5-m high cliff with a nearly 80o angle face (Fig.
7.3.4). The long axis of the cliff is oriented at an azimuth of approximately 80 o from North,
almost in E-W direction, with the face almost perpendicular to N-S. This would indicate
possibility of topographic effects in the N-S component of the AMNT record. Since the
earthquake propagated from North to South, we would expect small amplification because the
waves are traveling away from the slope face rather than into it (Ashford et al., 1997). Waves
traveling into the face of the slope tend to strike the slope causing an increase in the measured
ground motion. Waves traveling away from the face tend to pass by the cliff, reducing the
measured ground motion compared to that of a vertically propagating wave. Based on Ashford
et al. (1997), topographic amplification is expected in a frequency range of ft = Vsavg / 5 H,
where H is the height of the slope (estimated ~ 24.5 m) and Vsavg is the time-average Vs over
the height H (429 m/s, as per Vs on Fig. 7.3.7), resulting in approximate ft of 3.5 Hz (0.285 s).

Beyond topographic effects, soft soil site effects may have also contributed to amplification
of the ground motion. For the depth to bedrock of about 31.6 m below the surface and average
Vs of 439 m/s, the fundamental site period fs is ~ 3.47 Hz or 0.288 s, based on fs =Vsavg / 4H,
which is very close to the frequency ft from topographic effects in the N-S direction.

The Fourier amplitude spectrum of the AMNT recorded ground motion during the main
event shown on Fig. 7.3.9 indicates two frequencies with higher amplitudes: 2-3.5 Hz (0.5-
0.29 s) and 10 Hz (0.1 s). The higher, 2.0-3.5 Hz range, matches the larger cliff topographic ft.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-46
However, the E-W component has higher amplitude than the N-S one, which is not indicative
of topographic effects. The upper bound of this frequency range also matches the estimated 1D
site frequency of 3.5 Hz. Based on these observations, it is likely that the amplitude increases
in the 2.0-3.5 Hz range due to 1D amplification site effects, with minor contribution from
topographic effects. In the lower, 10-Hz range, the N-S component has higher amplitude than
the E-W one by about 60%, while both components exhibit a spike. This frequency is within
the range identified during the HVSR (Fig. 7.3.8), which was not supported by the surface
wave measurements that did not indicate such impedance contrast. This may be attributed to
spatial variability of the top layer thickness under the station compared to the thickness of this
layer under the surface wave arrays.

Figure 7.3.9. Manta AMNT station: Fourier amplitude spectrum of the recorded ground motion
during the main earthquake event.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-47
PORTOVIEJO

Overview and Damage

One of the most affected cities from the April 16th earthquake was Portoviejo. Rapid
damage assessment was performed by the Risk Management Secretariat COE3 (Emergency
Operations Committee, Comité de Operaciones de Emergencias) using ATC-20 procedures,
with results presented on Fig. 7.3.10 Comité de Operaciones de Emergencias (Emergency
Operations Committee). The damage was mainly concentrated in the downtown area, where
most of buildings taller than 4 stories are located. The damage variability based on the number
of floors is presented on Fig. 7.3.11 that clearly shows that the number of affected buildings
increases with height and their estimated fundamental structural period on different zones of
the alluvial clayey sand deposits (green area of Fig. 7.3.10), which may indicate site effects.

Figure 7.3.10. Portoviejo: Rapid assessment using ATC-20, with color tagging based on inspection
damage level (green: minor; yellow: partial; red: major). Map and data developed by EC COE3.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-48
The average estimate of damage of Fig. 7.3.11 was developed by first determining the
average estimated damage using the COE3 data Then, following the ATC procedure, the
average damage within a frequency range was multiplied by the frequency and divided by the
sum of frequencies to derive the average estimate of damage.

Figure 7.3.11. Portoviejo: Average estimated damage in Portoviejo City based on ATC procedures and
data from EC COE3.

Site Effects

The geologic map of Portoviejo is shown on Fig. 7.3.12 with blue circles indicating the
Ground Zero (most affected area) locations of APO1 station and Los Tamarindos with
characteristic soil profiles with energy-corrected Standard Penetration Test (SPT) N60 (in
blows/30cm) and shear wave velocity Vs values (in m/s) presented on Fig. 7.3.13. The
estimated elastic fundamental site periods Telastic (or Te) of 0.38 and 0.54 s for APO1 and Los
Tamarindos, respectively.

These profiles are representative of shallow soft clayey strata over rock, similar to those
evaluated in Finn and Ruz (2015) for the 2011 Tohoku earthquake. In this study, Finn and Ruz
presented a spectral amplification factor, expressed as the peak Spectral acceleration over the
Peak Ground Acceleration (Sa / PGA), of 4 to 5 for short periods between 0.1 and 0.5 s for
sites with VS30 from 200 to 300 m/s, which agree with the corresponding values at APO1 station
shown in Fig. 7.3.14.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-49
The elastic site period at APO1 Telastic of 0.38 s elongated to the inelastic value Tsite (or
Tinelastic) of 0.5 s based on the recorded peak Sa from the main event at this station. Nonlinear
effects through stiffness degradation can be expressed by the ratio of Tsite/Telastic equal to 1.32,
or 32% increase in the site period from its elastic value.

Figure 7.3.12. Portoviejo geomorphological map with Ground Zero sites of APO1 station and Los
Tamarindos shown as blue circles. Base map developed by Instituto Espacial Ecuatoriano (IEE) in
graphic scale 1:45,000.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-50
Vs : m/s

N60 : bl/30cm

Figure 7.3.13. Soil profiles of SPT N60 and Vs , and rock cores at APO1 and Los Tamarindos sites.

T (s)

Figure 7.3.14. Acceleration (Sa) and displacement spectra (Sd) spectra the record at APO1 station
during the main event and corresponding spectral amplification factor, Sa/PGA.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-51
GUAYAQUIL

Overview

The city of Guayaquil is the main seaport, on the left bank of the Guayas River, and largest
urban area of Ecuador, with a population of more than 2.3 million. It is heavily industrialized,
and its productivity represents over 20% of the country’s economy. The total area of Guayaquil
is about 345 (344.5 km2) split between land and water by 317 km2 and 28 km2, respectively.

Guayaquil is located on the left bank of the Guayas River along the S. American Pacific
coast. The Guayas River catchment of 34,500 km2 is among the largest in the Americas. Much
like most cities located at the edge of a navigable river, the underlying soils at Guayaquil were
deposited under water, and are hence are generally soft and compressible. Hard rock of the
Cretaceous period underlies the deep soft sediments (Reynaud et al., 1999). The ground is
mostly in low elevations with groundwater close to the surface, which often requires
compaction of the upper fills before construction.

Geotechnical Zones

The geotechnical zones of Guayaquil and the city’s strong motion stations are shown on
Fig. 7.3.16 after Vera-Grunauer (2014). The characteristics of each zone are described below.

Geotechnical Zone D1 to D3: Estuarine Deltaic Deposits


The Estuarine Deltaic Deposits are divided in 3 sectors: geotechnical Zone D1, which
corresponds to the estuarine-deltaic deposits in the central and southeast zone of Guayaquil
City; Zone D2, which corresponds to the southern zone of the city; and Zone D3 (subdivided
in D3a and D3b), which corresponds to the NE and SE zones of the city. Zone D3a corresponds
to deposits with elastic fundamental site periods Te < 1.6 s and D3b has Te > 1.6 s.

Regional clays are usually fragmented and have been homogenized into the silty clay
matrix by burrowing organisms. The often have diatoms in their composition, the amount of
which depends on the location and depth within the deltaic-estuarine zone. Microscopic studies
have revealed a sandy organic mud texture with abundant well-preserved diatoms and common
plant microfragments, with local cementation of the framboidal pyrite sediment. Zone D3
seems to have more pyrite concentration than Zone D1.

Geotechnical Zone D4: Alluvial Valley Deposits


The pyrite cementation present in the deltaic estuarine clays of Zones D1 to D3 is the main

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-52
difference with the alluvial clays of Zone D4 in northern Guayaquil. In this area, the Cayo
formation bedrock is deeper than in the deltaic-estuarine zones. Subzones depend on depth to
rock: D4a with depth < 10 m, D4a with depth of 10-20 m, and D4c with depth > 20 m.

Figure 7.3.14. Geotechnical Zones of Guayaquil (Vera-Grunauer, 2014).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-53
Geotechnical Zone D5: Alluvial-Lacustrine Deposits
In Zone D5, it is common to find a dark to greenish grey clay layer over the Cayo formation
with a thickness of 4-15 m. Colluvial-alluvial soils are found in riverbeds or at the base of the
hills, as can be seen in the ESPOL Prosperina campus and at Colinas de Los Ceibos, where
layers of expansive black clay of lacustrine origin with thickness of several m are found.

Geotechnical Zone D6: Colluvial Deposits


In the southern flanks, the colluvial soils can be few m thick (e.g., Cooperativa San Pedro).
These are typically red clays, hard when dry but very plastic when wet, with variable amounts
of siliceous shale blocks called “chert.” The El Paraíso urbanization lies partly over a dejection
cone that are often present in these flanks.

Geotechnical Zone D7: Residual Deposits and Rock Outcrop


A well-marked lithologic control is present in Zone D7. Soil overlying the Guayaquil rock
formation are typically brick-red and only 1-3 m thick. They become thicker in Durán hills and
in western Guayaquil, where overburden reaches a thickness of 30 m. Soils developed from
shales of the Cayo formation are red or reddish-yellow and are thicker than those developed
from sandstones. Around Jordán hill, thick brick-red soils have developed. This zone coincides
with a NE-SW trending fault whose presence must have favored this development.

Site Response of Estuarine Deltaic and Alluvial Deposits

Guayaquil is mostly built over estuarine-deltaic and alluvial deposits, with only small
portion in the north and northwest area built on shallow lacustrine and colluvium/residual
deposits. Figure 7.3.15 shows a typical geologic section in the downtown area.

The strong motions stations are shown on Fig. 7.3.14. Of particular interest is the Ramon
Unamuno Stadium (AGYE2) station, where the Institute of Geophysics (IGN) has installed a
broadband accelerometer. The instrument is in free-field conditions of deep soft soil deposits
(code-type F) with a Vs30 of 100 m/s. Geotechnical investigations in this area were performed
in 2005 and 2014 (details in Vera-Grunauer, 2014). The 2005 investigation included SPT with
selective sampling, CPT, dynamic CPT, DCP, vane shear, VST, and Vs measurements, SASW.
In 2014, a deep geotechnical boring, ERU, was drilled to approximately 145 m with SPT
performed until refusal (defined as N60 > 60 bl/30cm) at 37.5 m; below this depth a rotary
drilling was used to a depth of about 107 m. In addition, CPTu down to 39 m and MASW
combined with ReMi to derive Vs profile were performed at the site.
GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations
Report Version 1 7-54
Figure 7.3.15. Typical subsoil profile at downtown Guayaquil.

The soil profile at AGYE2 is a typical example of estuarine deltaic Guayaquil deposits,
comprised of 1.0 m granular fill which overlays a well-defined sequence of clay and sand strata.
In particular, the topmost clayey layers consist of 2.5 m of dark yellowish clay, underlain by
16.5 m of clay interlayered by thin seams of silty sand. Below this layer, 16 m of greenish gray
stiff clay are found, followed by 4 m of hard clay. The Plasticity Index, PI, of this clayey
deposit varies from 40 to 60 at the top 5 m, and from 20-40 up to 38 m. Below 40 m, there is
very dense sand layer, approximately 12 m thick. A thick clayey stratum is detected from 52
to 80 m, consisting of a greenish gray hard clay with unconfined compression strength, qu ≥
0.5 MPa. About 25 m of very dense sand underlays the clay, followed by a layer of greenish
gray hard clay with sporadic gravel. The groundwater table was detected at depth of 2.5 m
below the ground surface.

The Vs profile derived from data collected from the two subsurface investigations, either
directly with geophysical testing, or with correlations of the Vs or low-strain shear modulus G
with the SPT and CPT results are presented on Fig. 7.3.16.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-55
Vs (m/s)

PRES. HID.
(t / pulg )
2
COLOR
UNIFIED SOIL
0 200 400 600 800 1000 1200 1400 CLASSIFICATION SYSTEM

0
Estadio Ramón Unamuno, ERU 2014
CH
VsA = 120 (s'o/Patm)0.273 (Soft clay) ; Lin et al., 2014
Gmáx = CG3 Cub1 ex (so'/Pa)nG (Granular geomaterials) ; Menq, 2003 MH
CH
Vs = 280 (s'o/Patm)0.261 (Dense sand) ; Lin et al., 2014 SM
CH
SC
VsB = 230 (s'o/Patm)0.261 (Hard clay) ; Lin et al., 2014 CL
SM
CH
SM
Gmáx = Pat 380 e-1.3 (P/Pat)0.50 (Calibrated for GYE-clay) ; CH
modified of Pestana and Salvati, 2006
MH
Profile Selected
210 ECU (SASW at Puerto Azul, 2005)
20 216 ECU (SASW at Estadio Ramón Unamuno, 2005)

Young estuarine deltaic CH

Fine-grained soils
CL

SM

40
Poorly-graded SAND with SILT;
very dense;
dark gray

Dense Sand Poorly-graded SAND;


very dense;
dark gray

Poorly-graded SAND with SILT;


1.00 very dense; dark gray

SANDY CLAY; hard;


greenish gray

60
Hard Clay (qu>0.5MPa)
CLAY; hard;
greenish gray

SILT; grreenish gray


Depth (m)

CLAY; hard;
greenish gray

80
Well graded SAND with SILT;
very dense; dark gray

Very Dense Sand 1.50


Poorly graded SAND with SILT;
very dense; dark gray
Well graded SAND with SILT;
very dense; dark gray

100
SAND with SILT;
very dense; dark gray

GRAVEL with SAND and SILT;


4 cm; subrounded

CLAY: hard; greenish gray

Hard Clay (qu>0.5MPa) SILT; hard; greenish gray

120
CLAY: hard; greenish gray

GRAVEL with SAND and SILT;


4 cm; subrounded

Poorly graded SAND with SILT;


very dense; dark gray
GRAVEL with SAND and SILT;
4 cm; subrounded

Soft Claystone
SAND with SILT; dark gray
1.75

Soft CLAYSTONE

140

HALFSPACE, Vs=2800 m/s

Figure 7.3.16. Geotechnical description and Vs profile at station AGYE2.

The site response at the AGYE2 soft and deep site was evaluated using as input 3 recorded
ground motions, 2 of which have short epicentral distance and 1 is from far field. The first
ground motion is from the October 28, 2012 Mw4.9 earthquake which occurred at an
approximate distance of 9 km from AGYE2 (or boring ERU). The recorded PGA was 0.088 g
in the EW direction and 0.05 g in the NS direction. The second ground motion occurred 70 km

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-56
away from Guayaquil on April 28, 2015 with Mw of 5.9 and PGA of 0.086 g (EW) and 0.13 g
(NS). Records from the main April 16, 2016 Mw7.8 event were as the third event, with an
epicentral distance of 170 km and a PGA of 0.098 g (NS) and 0.094 g (EW).

Although the maximum PGA values in the 3 earthquakes are of the same order of
magnitude, the predominant high frequency content of the near motions did not resonate with
the ERU site that has a low fundamental frequency. In contrast, the distant subduction 2016
event with low predominant frequency, did resonate with the site, as shown in the response
spectra of Fig. 7.3.19. The same figure indicates the inelastic fundamental site period is Ts of
1.67 s, which results to a ratio with the elastic site period of Ts/Telastic of roughly 1.08.

The distant subduction 2016 earthquake event generated a peak Spectral Acceleration Sa
that is 15 times higher at AGYE2 (VS30 = 100 m/s) than at AGYE station that lies on intrusive
igneous rock (VS30 = 1,800 m/s) at the inelastic site period (1.67 s). The second fundamental
site period of ~0.55 s was also strongly amplified by this event (Fig. 7.3.19), a phenomenon
also observed in similar profiles in Eastern United States (Nikolaou et al., 2012).

Station AGYE2
Vs30 = 101 m/s
Site Type F (Deep soft soils)
Sa (g)

larger

T (s)

Station AGYE
15 x

Vs30 = 1800 m/s


Site Type A (Hard rock)
Sa (g)

T (s)
Figure 7.3.17. Response spectra obtained for AGYE
Measured andNorth
GYE-ERU AGYE2 records
Component from three EC earthquakes.
(4/16/16)
Measured GYE-ERU East Component (4/16/16)
GEER-ATC Muisne, Ecuador, Earthquake Measured GYE-ERU North Component (4/28/15)
7-Geotechnical Observations
Report Version 1 Measured GYE-ERU East Component (4/28/15) 7-57
Measured GYE-ERU North Component (10/28/12)
Measured GYE-ERU East Component (10/28/12)
Finally, based on deconvolution analyses in Vera-Grunauer (2014), Fig. 7.3.18 shows a
comparison of the response of the ERU profile when subjected to the previous 3 ground
motions and 2 additional local ground motions recorded at AGYE2 in 2012, with
characteristics of moment magnitude Mw and distance of the recording station AGYE2 from
the epicenter, R, shown on Table 7.3.2.

Table 7.3.2. Characteristics of five local seismic events recorded at AGYE2, near Boring ERU.
h Component (4/16/16)
Component (4/16/16) Station ERU R (km) Mw
h Component (4/28/15)
Event 1 (4/16/16) 286 7.8
Component (4/28/15)
h Component (07/30/12) Event 2 (4/28/15) 12.5 5.8
Component (07/30/12)
Event 3 (07/30/12) 22.9 4.9
h Component (10/28/12)
Component (10/28/12) Event 4 (10/28/12) 101.4 5.3
h Component (06/27/12)
Component (06/27/12)
Event 5 (06/27/12) 136 5.2

The site response is presented on Fig. 7.3.18 in terms of lateral displacement, maximum
shear strain , ratios of post-shaking over low strain shear modulus G/Gmax, and shear wave
velocity Vs/Vs,max, Cyclic Stress Ratio, CSR (taken as 0.65 of the cyclic shear / effective vertical
stress) and Peak Ground Acceleration (PGA) with depth. From the 5 ground motions examined,
the April 2016 event was the one generally felt most at the site, especially at shallow depths,
due to the resonance effects explained earlier. Accelerations from the April 2015 event was
only slightly higher at the surface and higher at larger depths, which did not cause significant
consequences in the overall site response at the ground surface or observed damage.
Lateral Displacement Max Shear Strain Peak Acceleration
(m) (%)
G / Gmax Vs / Vsmax CSR (g)
Lateral Displacement Max Shear Strain Peak Acceleration
(m) (%)
G / Gmax Vs / Vsmax CSR (g)
Depth (m)
Depth (m)

GYE-ERU North Component (4/16/16)


GYE-ERU East Component (4/16/16) Station ERU R (km) Mw
GYE-ERU North Component (4/28/15)
Event 1 (4/16/16) 286 7.8
GYE-ERU East Component (4/28/15)
GYE-ERU North Component (07/30/12) Event 2 (4/28/15) 12.5 5.8
GYE-ERU East Component (07/30/12)
Event 3 (07/30/12) 22.9 4.9
GYE-ERU North Component (10/28/12)
GYE-ERU East Component (10/28/12) Event 4 (10/28/12) 101.4 5.3
GYE-ERU North Component (06/27/12)
GYE-ERU East Component (06/27/12)
Event 5 (06/27/12) 136 5.2

Figure 7.3.18. Response of ERU profile subjected to the 5 local ground motions of Table 7.3.2 in terms
of lateral displacement,
GYE-ERU Northmax shear
Component strain , ratios of G/Gmax, Vs Station
(4/16/16)
GYE-ERU East Component (4/16/16)
/Vs,maxERU
, CSR, Peak Ground
R (km)
Acceleration,
Mw
PGA as functions of depths
GYE-ERU below
North Component the ground surface.
(4/28/15)
Event 1 (4/16/16) 286 7.8
GYE-ERU East Component (4/28/15)
GYE-ERU North Component (07/30/12) Event 2 (4/28/15) 12.5 5.8
GYE-ERU East Component (07/30/12)
Event 3 (07/30/12) 22.9 4.9
GEER-ATC Muisne, GYE-ERU
Ecuador,North Component (10/28/12)
Earthquake 7-Geotechnical Observations
GYE-ERU East Component (10/28/12) Event 4 (10/28/12) 101.4 5.3
Report Version 1 GYE-ERU North Component (06/27/12) 7-58
GYE-ERU East Component (06/27/12)
Event 5 (06/27/12) 136 5.2
7.4 Liquefaction
INTRODUCTION

Liquefaction was observed throughout the affected areas due to the presence of loose soils
and shallow groundwater table. Scattered signs of liquefaction were detected within the entire
province of Manabí. One of the most extensive liquefaction manifestation was observed in the
Port of Manta by the GEER-ATC reconnaissance team that spent substantial part of our stay
to map liquefaction induced phenomena and performed in-situ testing for further studies. Some
cases of liquefaction were identified in the city of Manta and along the Portoviejo river. Two
macro-landslides in Manabí, discussed in Section 7.6, were likely triggered by increment in
water pressure and reduction in soil strength due to potential liquefaction that was not however
evident on the surface during reconnaissance in the immediate vicinity of the landslides.

This section briefly discusses some definite cases of liquefaction manifestation, and mainly
focuses on the Port of Manta liquefaction and earthquake-induced settlements. The next
version of this report will contain additional information being processed at this time. Observed
settlements, with or without evidence of liquefaction, are discussed throughout this Chapter.

LIQUEFACTION MANIFESTATION OBSERVATIONS

Portoviejo River

Liquefaction signs were detected during the reconnaissance army helicopter flyover in the
route on Fig. 7.4.1 between Portoviejo and the Poza Honda Dam (see Chapter 2) with sand
ejecta photos taken along Rio Portoviejo presented on Fig. 7.4.2.

POZA HONDA DAM

FLIGHT ROUTE

Figure 7.4.1. Flyover route along Portoviejo River were liquefaction was observed.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-59
Liquefaction signs were detected during the reconnaissance army helicopter flyover in the
route on Fig. 7.4.1 between Portoviejo and the Poza Honda Dam (see Chapter 2) with sand
ejecta photos taken along Rio Portoviejo presented on Fig. 7.4.2.

Figure 7.4.2. Liquefaction evidence along Portoviejo River. Photos from reconnaissance army
helicopter flyover (coordinates pending next version).

Mehia Bridge

The approach embankments of the Mejia Bridge, on the road between Portoviejo and
Crucita, experienced extensive vertical settlement and lateral deformations due to instability
that may be indicative of liquefaction, although surface manifestation of sand ejecta was not
identified in reconnaissance. Details on this failure are presented in Embankments Section 7.9.

An aerial view of the failure at the Mejia Bridge site (by COE-3) is presented on Fig. 7.4.3
with sketched markups of the observed embankment movement. Gabion slopes were
constructed as countermeasures upstream of the abutments along both river banks. The gabions
on the eastern abutment failed either due to sliding or rotation. The eastbound bridge moved
about 20 cm, separating from the westbound structure, while the west side performed well.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-60
The assumed failure planes on Fig. 7.4.3 appear to be either along the interface between
the embankment and the foundation soils (planar surface), or through the foundation soils.
The latter, indicative of a more circular surface failure may be due to liquefaction-induced
softening that allowed the reduced-strength soil to shear during the earthquake.

Figure 7.4.3. Mejia Bridge: Aerial view of embankment failure with assumed movement and failure
modes marked with yellow (after COE-3, GPS: 0°59'24"S, 80°28'11"W).

Boca de Briceno Bridge

The Boca de Briceno Bridge, discussed in Section 7.7, has subsurface soil conditions that
consist mainly of loose silty sand to sandy silt in top 2 to 5 m with shallow ground water table
(GWT) found at or within 0.6 m below the ground surface (see Section 7.2).

Geopiers were used on the original design as a ground improvement mitigation measure to
address potential liquefaction of the loose silty sands, and potential stability and settlement
concerns for the embankment. It is estimated that this site experienced approximately the
design acceleration during the main April 16th event, with liquefaction evident in a free-field
area adjacent to the embankment, as shown on Fig. 7.4.4 with examples of sand boils and sand
ejecta through cracks. The bottom right photo of Fig. 7.4.4 shows settlement of the ground
around the abutment footing. In contrast to other embankments in its vicinity this embankment
performed well, likely a beneficial result of the ground improvement.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-61
Figure 7.4.4. Boca de Briceno Bridge liquefaction evidence in free-field and adjacent to embankment
footing (top, est. GPS: 0°30'59"S, 80°26'31"W; bot. left, GPS: 0°30'59.1"S, 80°26'29.7"W, bot. right,
GPS: 0°30'58.98"S, 80°26'28.93"W).

Manta Overpass

The Manta overpass site experienced strong ground shaking and surface evidence of
liquefaction was present at several locations as shown on Fig. 7.4.5. There was also severe
surface cracking of the pavement and sidewalk near a Mobil gas station and ground settlement
around the underground gas tanks. As discussed in detail in Section 7.2, a post-earthquake
geotechnical investigation showed a loose silty sand layer with low plasticity fines in the top
1.5-3 m with the water table within a depth of 2 m from the existing ground surface. This layer
may have liquefied during the main event which could have resulted in the cracking and ground
deformation that observed near this area.

This case is discussed extensively in Section 7.7, where more photos of surface cracking
are presented.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-62
Figure 7.4.5. Manta overpass: Possible evidence of sand boils and cracking due to lateral movement
near the site (GPS: 0 o57'1.7"S, 80o43'8.7"W).

Tarqui Parish, Manta

The parish of Tarqui is part of the Manta canton in Manabí province. Multiple sites showed
evidence of liquefaction as shown on Figs. 7.4.6 and 7.4.7 that display typical manifestation
observations with sand ejecta. Locations where liquefaction was observed in Tarqui are
identified in the COE-3 maps of Fig. 7.4.8.

Figure 7.4.6. Signs of liquefaction in Tarqui, Manta (GPS: 0 o57'9.1''S, 80o42'29.4''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-63
Figure 7.4.7. Signs of liquefaction in Tarqui, Manta (GPS: left, 0 o57'20.0''S, 80o42'39.9''W; right,
0o57'20.4'', 80o42'38.4''W).

Figure 7.4.8. Liquefaction locations identified by COE-3 in Tarqui, Manta.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-64
Along the Coast of Manabí

Potential signs of liquefaction were detected during the GEER-ATC army helicopter
flyover along the coast of Manabí with sample of photos on Fig. 7.4.9. The land appeared
settled and flooded and areas where sand ejecta seemed to be present are circled in the photos.

Figure 7.4.9. Liquefaction evidence from helicopter flyover of Manabí coast (GPS clockwise, top left:
0o41'29.2''S, 80o15'10.7''W; 0o39'58.2''S, 80o18'17.2''W; 0o39'41.8''S, 80o18'46.6''W; 0o12'31.7''S,
80o15'22.5''W; 0o9'37.9''S, 80o15'47.2''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-65
PORT OF MANTA

Liquefaction Observations

Manta is one of the five major ports in Ecuador. Shown on Fig. 7.4.10, the port facility
consists of a pile-supported wharf, two pile-supported piers, a rock-fill breakwater. A parking
storage area for new imported cars, shown on Fig. 7.4.9, was constructed by placing backfill
material behind a rock berm. Liquefaction and liquefaction-induced phenomena were evident
throughout the port. In addition, earthquake-induced damage and displacement of the pile-
supported structures was observed even without clear manifestation of liquefaction.

Rock breakwater
& embankment

Pile-supported
piers

Pile wharf

Rock berm
with backfill

Figure 7.4.10. Aerial view of Manta Port with main geo-components (Digital Globe, ©2016).

At the parking storage area of the Manta Port (Fig. 7.4.11), liquefaction-induced lateral
spreading towards the rock berm produced a number of cracks parallel to the rock berm along
with ejecta. In addition, large sand boils erupting from a single point developed at several
locations. The soil type of the ejecta ranged from sand to silt at various locations and samples
were obtained for subsequent laboratory testing.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-66
A

Rock berm

Boat ramp,
RC retaining walls

Figure 7.4.11. Manta parking area: Aerial view of backfill behind rock berm (Digital Globe, ©2016).
Settlement and lateral displacement were measured by the reconnaissance team along the orange lines.

Figure 7.4.12 shows one of the large soil boils that erupted primarily from a single vent in
the pavement. Eyewitness accounts indicate that water sprayed out of the ground as high as 1
m immediately following the ground shaking. In contrast, Fig. 7.4.13 shows sand ejecta that
emanated from transverse cracks in the pavement as a result of lateral spreading.

Liquefaction and lateral spreading caused foundations for light poles shown in Fig 7.4.14
to rotate and sink into the ground. Rotation was approximately 2º towards the rock berm and
1.5º degrees to the west. Vertical wide cracks were observed along the perpendicular non-
retaining walls of the parking as shown on the bottom right of Fig. 7.4.14.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-67
Figure 7.4.12. Manta Port parking area: concentrated sand blow in the center of pavement
observed by GEER co-leader S. Nikolaou (GPS: 0°56'27.6"S, 80°43'29.4"W).

Figure 7.4.13. Manta Port parking area: Sand vents erupting from cracks in pavement formed by lateral
spreading. (left: 0°56'26.9"S, 80°43'29.5"W; right: 0°56'29.5"S, 80°43'27.8"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-68
Figure 7.4.14. Manta parking area: Vertical offset in ground retaining wall, wall deformation and light
pole rotation (left). Wide crack at the perpendicular (non-retaining) wall next to GEER team member
K. Rollins (right). Pole coordinates from GPS (0°56'26.7"S, 80°43'29.8"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-69
A boat ramp on the southeast side of the area (see Fig. 7.4.11) with 4-m high reinforced
concrete retaining walls moved about 40 cm outward at the top of the wall as shown in Fig.
7.4.15 (top). This led to significant vertical cracks in the walls on either side of the boat ramp.
In addition, the floor slab at the base of the boat ramp heaved upwards and developed
significant longitudinal cracks (see Fig. 7.4.15, bottom).

To document the lateral spreading displacements that occurred in this area, crack widths
along 5 lines normal to the rock berm were measured as shown on Fig. 7.4.11. In addition, the
distance to each crack was measured relatively to the face of a concrete panel wall running
roughly parallel to the top of the rock berm.

Figure 7.4.15. Lateral movement of 40 cm at the 4-m tall retaining walls adjacent to boat ramp (top)
and heave and longitudinal cracking of base slab (bottom, GPS: 0°56'28.9"S, 80°43'26.8"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-70
Figure 7.4.16 shows cumulative lateral displacement up to the wall, which is considered
the reference 0-point for each line. The average lateral displacement was about 37 cm, with a
range of 8 to 70 cm. Lateral displacement increased substantially within ~10 m of the wall,
approaching the edge of the rock berm slope, and was < 2 cm beyond 90 m behind the wall.
The measured variation in the lateral displacements could be useful in evaluating the effect of
subsurface stratigraphy and soil properties in lateral spreading displacement prediction models.
80

70
Cumulative Lateral Ground Displacement (cm)

Line 1
60
Line 2

50 Line 3
Line 4
40 Line 5

30

20

10

0
0 20 40 60 80 100 120 140
Distance from the Wall Face (cm)
Figure 7.4.16. Manta Port parking area: Cumulative lateral spread displacements along the 5 lines
perpendicular to the shoreline shown on Fig. 4.7.11.

Vertical offset in the pavement, shown on Fig. 7.4.17, was observed along the cyan Line
A-A of Fig. 4.7.11 at a distance of about 3.5 m landward at 5-m intervals. Typically, the ground
settled downwards on the landward relative to the seaward side, with maximum offsets of 50
and -14 cm with average of 15 cm.
60
50
Vertical Offset (cm)

40
30
20
10
0
-10
-20
0 50 100 150 200 250 300
Distance from Northwest Side (m)

Figure 7.4.17. Manta Port parking area: Vertical offset in concrete pavement produced by lateral
spreading along the cyan Line A-A of Fig. 4.7.11, at 5-m intervals. Line A-A is roughly parallel to the
concrete wall at a distance of about 3.5 m from the wall face.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-71
Performance of Piles: Piers and Wharf

The reconnaissance team observed the Manta Port wharf structures both by land and by
motor boats (Fig. 7.4.18), arranged by our Ecuadorian colleagues and the Port Authority,
giving us an in-depth view of the wharf performance and the soil-pile interaction behavior.

Figure 7.4.18. Manta Port wharfs: Reconnaissance team on motor boats (GPS 0°56'16.6"S,
80°43'29.3"W).

The piles supporting the two piers at the northern end of the port showed little damage.
These piers are supported by multiple rows of square reinforced concrete (RC) piles, about 0.5-
m wide, with 15-20 piles in each row as shown in Fig. 7.4.19a. Most of these piles were driven
vertically with some batters at the ends of each row, likely to increase lateral restraint. In
interior pile rows, shear cracks in beams used to attach an additional pile to the pile cap were
observed (7.4.19b). At the NW corner of the southern pier, some heads of battered and square
piles experienced cracks and compression failure (Fig. 7.4.20). A nearby CPT showed the top
7.5 m of soil is a potentially liquefiable sand underlain by dense soils that caused cone refusal.
In contrast to the generally good performance of the piles supporting the piers, the piles at the
wharf parallel to the shoreline (Fig. 7.4.20) experienced considerable damage.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-72
Shear
cracking

(a) (b)

Figure 7.4.19. Manta Port wharf: (a) Rows of piles supporting the two piers on the northern end of the
port and (b) shear cracks in beam connecting a pile to the pile cap (GPS: 0°55'58.2"S, 80°43'18.8"W).

(a) (b)

Figure 7.4.20. Manta Port wharf: (a) cracking and (b) compression failure of piles at the northwest
corner of the southern pier (GPS: 0°55'58.2"S, 80°43'18.8"W).

Typically, the 0.4-m square RC piles on the seaward side of the wharf were battered at a
1H:5V slope and spaced 3 m OC (on center) along the wharf, with alternate forward/reverse
direction of batter (Fig. 7.4.21). After the earthquake, a crack with vertical and horizontal
offsets relatively to the ground developed at the west side of the pier.

The offsets were measured by our team along the yellow line of Fig. 7.4.22 that also
indicates intervals of 100 m along the wharf for reference. In the first 100 m, offsets were small
and pile damage minimal. Between 100-300 m, the vertical offset increased to about 14 cm on
average, with pile damage still relatively small. However, in the zone from 300 m onward, the
vertical offset increased to ~40 cm and with an increase on the horizontal as well. Shown on

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-73
7.4.21, the offsets escalation associate well to the observed damage to the piles supporting the
structure shown on Figs. 7.4.24 through 7.4.26.

Figure 7.4.21. Manta Port wharf: Battered piles at 3-m OC supporting seaward side of the wharf with
alternate sloped 1H:5V batters with rock riprap slope behind them (GPS: 0°56'22.7"S, 80°43'33.8"W).

As shown on Fig. 7.4.23, the piles with a reverse batter exhibited considerable damage and
offset at the pile head while piles with a forward batter did not. Port engineers indicated that
many of the piles had been retrofitted to improve the strength of the pile-to-pile cap connection.
Unfortunately, the retrofit was not generally successful and reinforcing bars pulled out or
sheared off at the pile head. For a given lateral deflection towards the sea, the piles with a
reverse batter, would be in tension while the piles with a forward batter would be expected to
be in compression. It is likely that this difference in loading is associated with the observed
damage pattern. Photos in Fig. 7.4.21 show that there was a significant number of cases where
the piles with reverse batter moved 30 to 40 cm beyond their original connection location on
the deck after shearing off from the deck. There is no visible indication of cracking or lateral
spreading in the gravel berm behind the piles, although liquefaction and lateral spreading may
have occurred in the immediate vicinity of the piles.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-74
Damage to
reverse and forward
battered piles

Damage to
reverse battered piles

No pile damage

Figure 7.4.22. Manta Port wharf: Aerial view of pile-supported wharf along shoreline and areas with
pile damage. Offsets measured along the yellow line (GPS: 0°56'12.4"S, 80°43'23.5"W).

60
Vertical Displacement
Horizontal Displacement
50

40
Displacement (cm)

30

20

10

0
0 100 200 300 400 500 600 700
Distance (meters)
Figure 7.4.23. Manta Port wharf: Measured vertical and horizontal offset as a function of distance along
the yellow line parallel to the shore in Fig. 7.4.22.

At some locations near the end of the wharf (Fig. 7.4.22), damage was observed at the pile
head for both the piles with forward and reverse batter as shown in Fig. 7.4.26. The inclination
of the piles with reverse batter seems somewhat less in the photos and the vertical offset was
greatest towards the end of the pier.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-75
Figure 7.4.24. Manta Port wharf: Cracking and shearing of pile head connection for backward-batter
piles, with minimal damage to forward-batter piles (GPS: 0°56'14.42"S, 80°43'27.41"W).

Figure 7.4.25. Manta Port wharf: Failure of pile head in backward-batter piles, with 30-40 cm
horizontal offset. Forward-batter piles were not damaged (GPS: 0°56'13.6"S, 80°43'26.6"W).

Figure 7.4.26. Manta Port wharf: Distress and cracking of pile head connection for piles with both
forward and backward batter (GPS: 0°56'11.2"S, 80°43'24.8"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-76
Performance of Piles: Port Authority Operation Building

A 3-story building that houses operation offices of the Port Authority is located within the
Manta Port’s embankment between two of the piers, as shown in Fig. 7.4.27. At the time of the
reconnaissance team visit, repair work was being done at the zone of the embankment adjacent
to the building as shown in Fig. 7.4.28. The building is founded on piles that go through the
slope of the embankment possibly to a firm soil layer beneath the foundation soils of the
embankment, with the upper layer being loose and potentially liquefiable. The foundation
consists of rectangular RC piles with a width of 0.4 m.

Manta Port Authority


Operations Building

Figure 7.4.27. Manta Port Authority Operations Building: Location on aerial photo of the port (GPS:
0°55'56.3"S, 80°43'20.3"W).

Figure 7.4.28. Manta Port Authority Operations Building: Repairs performed in the embankment near
the building at the time of reconnaissance (GPS: 0°55'56.3"S, 80°43'20.3"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-77
The embankment spread laterally towards the building, resulting in measured lateral
displacement of 0.4 m and equal vertical settlement relative to the pile-supported structure as
shown on Fig. 7.4.29a,b,c. Tilt measurements indicate that the piles were tilting at angles of 4
to 6o at their top. The pile heads experienced some damage to the concrete as can be seen in
Fig. 7.4.24d.

40 cm

(a) (b)

40 cm

(c) (d)

Figure 7.4.29. Manta Port Authority Operations Building: (a) horizontal and vertical displacements of
embankment relative to building; (b) horizontal movement between building and embankment; (c)
vertical settlement of embankment; (d) damage at the pile head (GPS: 0°55'55.2"S, 80°43'19.6"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-78
7.5 Earth Retaining Structures
INTRODUCTION

The reconnaissance team observed seven retaining walls and masonry-type structures in
the Manabí province shown on Fig. 7.5.1, that experienced from little or no damage to poor
performance or failure. The retaining systems observed were mostly rigid, Cast in Place (CIP),
cantilever walls and masonry structures. Mechanically Stabilized Earth (MSE) walls were not
identified, as this technology is not used extensively in Ecuador. This section provides detailed
reconnaissance information on the observed retaining walls and systems listed on Table 7.5.1.

Figure 7.5.1. Locations of 7 retaining walls and masonry-type structures observed in reconnaissance.
Table 7.5.1. Retaining structures details and damage. Actual wall heights could not be confirmed as
they were partially buried; for Mejia Bridge, heights are based on design drawings.

Location in Manabi Province Retaining Wall Characteristics


Name Damage
Latitude Longitude System max
Town Construction
(South) (West) Observed Height
Avenida 4 de Abutments (2) 2.5
Noviembre over Rio Manta 0⁰ 57' 11.26'' S 80⁰ 43' 4.19'' W Extensive repairs required
Manta Bridge Wing Walls (4) 6.0
Abutments (2) 3.5
Rio Chico Bridge Rio Chico 0⁰ 58' 35.65'' S 80⁰ 25' 21.29'' W Minor repairs required
Wing Walls (4) 5.0
Abutments (2) 5.3
Mejia Bridge Mejia 0⁰ 59' 22.73'' S 80⁰ 28' 11.14'' W Cast in Place Extensive repairs required
Wing Walls (4) 8.3

Boca de Abutments (2) 6.0


Boca de Briceno Bridge 0⁰ 30' 58.27'' S 80⁰ 26' 29.08'' W No damage observed
Briceno Wing Walls (4) 9.0

Bahía de Abutment (1) 3.0


Bahía (Las Caras) Bridge 0⁰ 36' 37.2'' S 80⁰ 25' 29.35'' W Minor repairs required
Caráquez Wing Walls (2) 4.5

Bahía Wall along Bahía de Unreinforced Structural failure


0⁰ 36' 30.61'' S 80⁰ 25' 25.03'' W Wall (1) 2.0
Avenida Virgilio Ratti Caráquez Stone Masonry Replacement required
Geotechnical, structural
Geotech, Structural
failure failure
Port of Manta Manta 0⁰ 56' 28.86'' S 80⁰ 43' 26.69'' W Cast in Place Wall (1) 3.0
Replacement required

Unreinforced Geotechnical failure


Velboni Supermarket Portoviejo 1⁰ 3' 37.78'' S 80⁰ 27' 29.27'' W Wall (1) 1.0
Brick Masonry Replacement required

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-79
AVENIDA 4 DE NOVIEMBRE BRIDGE ABUTMENT AND WING WALLS

The abutment walls and wing wall of the Avenida 4 de Noviembre Bridge over the Rio
Manta (0°57'11.3''S, 80°43'4.2''W) performed fairly well, with needs for repairs only to some
components and portions of the bridge itself. The bridge lies on bearings on top of the abutment
walls with the abutment walls and wing wall likely supported on deep foundations.

Typical damage observed at the east abutment and wing wall is shown on Fig. 7.5.2. The
damage appears to have occurred as a result of impact of the superstructure into the wing walls
causing them to fail in shear. There is also damage at the top of the abutment wall that may be
a result of horizontal movement of the bridge in the longitudinal direction, which could have
caused the bearing to react through friction against the top of the abutment wall.

(a) (b) (c) (d)

(e) (f)

Figure 7.5.2. Avenida 4 de Noviembre Bridge: Damage to east abutment and wing wall (GPS: (a)
0°57'10.1''S, 80°43'5.5''W, (b) 0°57'10.2''S, 80°43'5.6''W, (c) 0°57'10.4''S, 80°43'5.1''W, (d)
0°57'10.1''S, 80°43'5.2''W, (e) 0°57'9.8''S, 80°43'04.8''W, (f) 0°57'9.7''S, 80°43'4.6''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-80
The damage on the west abutment and wing walls is similar to the one at the east. At the
time of the reconnaissance, repairs were underway (Fig. 7.5.3), with cracked concrete being
removed exposing the reinforcing steel. Concrete will be cast to replace the damaged sections.
The wing walls away from the abutments had no observable damage.

Figure 7.5.3. Avenida 4 de Nov. Bridge repairs on west abutment and wing wall: cracked concrete is
removed, exposing reinforcing steel (left: 0°57'15.2''S, 80°43'2.1''W; right: 0°57'15.2''S, 80°43'2.2''W).

RIO CHICO (LAS CHACRAS) BRIDGE ABUTMENT AND WING WALLS

The Rio Chico (Las Chacras) Bridge is supported on a Cast in Place (CIP) concrete
abutment which. Based on recommendations in the 2013 geotechnical report by LUP this
bridge should be supported on driven piles, with the subsurface conditions described as various
layers of low plasticity clays, silts, and sands. Single digit Standard Penetration Test N-values
from 2 to 9 bl/30cm in the top 7 m, increased to a range from 20 to 50 bl/30cm up to a depth
of 32 m below the ground surface. The approach embankments to this bridge had significant
settlement and lateral displacement, as described in the embankments Section 7.9 of this report.

Although displacements of the approach embankments and lateral soil movement at the toe
of the abutment were observed (Fig. 7.5.4, right), the abutment and wing walls appear to have
performed well overall (Fig. 7.5.4, left), with needs for minor repairs primarily due to joint
movement between the superstructure and the approach embankment, as shown in Fig. 7.5.5.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-81
Deck

Wall
Toe slope
displacement
~30 cm

Figure 7.5.4. Rio Chico (Las Chacras) Bridge: East abutments and wing walls (left, 0⁰58'35.6''S,
80⁰25'21.25''W); Displacement of toe slope at east abutment (right 0⁰58'34.96''S, 80⁰25'21.4''W).

Figure 7.5.5. Rio Chico (Las Chacras) Bridge: joint damage between approach and superstructure
(GPS: 0°58'35.6''S, 80°25'21.3''W). Team member D. Alzamora on the right.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-82
MEJIA BRIDGE ABUTMENTS AND WING WALLS

Information on the Mejia Bridge in the Manabí province (0⁰59'22.7''S, 80⁰28'11.1''W) was
available from design drawings of 2011 for the improvements at km 32 of the Portoviejo-
Crucita Highway, from the Ecuador Ministerio de Transporte y Obras Publicas (EMTOP,
2011). Based on this information, this bridge is supported on Cast in Place (CIP) concrete
abutments founded on driven piles (Fig. 7.5.6). Borings by Nuques y Luque Ingenieros
Consultores (2001) indicate that the soil profile consists of a soft layered silt and clay layer to
a depth of 18 m continuing with stiffer layers of silts and clays to a depth of 58 m.

Figure 7.5.6. Mejia Bridge: Design drawing from improvements at the Portoviejo-Crucita Highway
from the Ecuador Ministerio de Transporte y Obras Publicas (EMTOP, 2011).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-83
The approach embankments experienced large vertical settlements and lateral deformations
that are discussed in the Embankments Section 7.9. Gabion slopes were constructed as
countermeasures upstream of the abutments along both river banks. The gabions on the eastern
abutment failed either due to sliding or rotation potentially attributed to liquefaction in this
area. In general, the embankments and gabions on the western side performed well (Fig. 7.5.7).

~ 20 cm

Figure 7.5.7. Mejia Bridge: West abutment, gabions, wing wall without damage (top, 0°59'24.4''S,
80°28'11.5''W); Damage to east abutment, wing wall, and gabions (bot., 0°59'23.3''S, 80°28'11.9''W).

Overall the abutments and wing walls of the Mejia Bridge performed well, especially on
the western side. Some damage was observed at the eastern side (Fig. 7.5.8), where the
abutment wall moved downward by 15 cm at the corner and rotated slightly by 1 to 2o towards
the river. The abutment at the center section between the eastbound and westbound bridges had
also moved downward by about 20 cm toward the south. The eastbound bridge had moved
about 20 cm, separating from the westbound structure. This movement may be an indication
of slope instability or a combination of instability and liquefaction-induced effects. The
presence of the abutment on the east side may have prevented greater movement of the
abutment towards the river, potentially causing more severe damage to the structure, which
may have induced increased downdrag and lateral load to the pile foundation.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-84
20 cm
15 cm

~ 20 cm
1-2 deg

Figure 7.5.8. Mejia Bridge: Lateral displacements at east abutments (left: 0°59'23.7''S, 80°28'10.3''W;
right: 0°59'26.1''S, 80°28'09.7''W).

BOCA DE BRICENO BRIDGE ABUTMENTS AND WING WALLS

The Boca de Briceno Bridge (0⁰30'58.3''S, 80⁰26'29.1''W), shown on Fig. 7.5.9, is


supported on CIP abutments and wing walls founded on driven piles, based on design drawings
by the Ecuadorian Ministerio de Obras Publicas (EMOP, 2010). The retaining walls did not
exhibit visible signs of distress (Fig. 7.5.10), while the approaches had minor settlement. The
bridge joints have either closed or opened indicating bridge movement. Some damage of the
concrete sidewalk along the bridge roadway was observed as shown on Fig. 7.5.11. Ground
improvement with stone columns was utilized at the approach embankments (see section 7.7).

Figure 7.5.9. Boca de Briceno Bridge aerial photo (GPS: 0°31'3.7''S, 80°26'43.8''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-85
(a) (b)

(c)

Figure 7.5.10. Boca de Briceno Bridge: (a; b) Abutment and wing wall (GPS 0°30'58.4''S,
80°26'29.7''W; 0°30'57.7''S, 80°26'28.9''W); (c) Approach embankment and bridge pavement looking
south (GPS 0°30'58.1''S, 80°26'29.2''W). No visible damage observed.

(b)

(a) (c)

Figure 7.5.11. Boca de Briceno Bridge: Damage to bridge joints and sidewalk (GPS: (a) 0°30'57.0''S,
80°26'29.7''W, (b) 0°30'58.5''S, 80°26'29.6''W, (c) 0°30'57.3''S, 80°26'30.2''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-86
BAHIA BRIDGE WEST ABUTMENTS AND WING WALLS

The Bahia Bridge (0⁰36'37.2''S, 80⁰25'29.4''W) has CIP abutments and wing walls. A cover
wall did not allow the reconnaissance team to inspect the abutment face. Based on the structure
type and location we assume deep foundations support the abutments. The west abutment
includes wing walls to contain the approach embankment (Fig. 7.5.12). The wing walls have 2
rows of bolts to support the walls (the upper row has one and the lower row two bolts), likely
part of the original construction (Fig. 7.5.13). Damage was observed at the intersection of the
abutment and north wing wall where a formed crack caused a 5-cm thick concrete piece to
spall off the wing wall (Fig. 7.5.14), which could allow water runoff into the embankment.

see Fig. 7.5.14

Figure 7.5.12. Bahia Bridge: Abutment and wing walls showing only minor damage to one of the
abutment walls (left, 0°36'37.2''S, 80°25'29.4''; right 0°36'37.7''S, 80°25'29.0''W).

Bolted
connections

Figure 7.5.13. Bahia Bridge: Bolted connections of wing walls (left 0°36'37.5''S, 80°25'29.7''W; right
0°36'37.5''S, 80°25'29.7''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-87
~ 5 cm

Figure 7.5.14. Bahia Bridge: Damage at intersection of abutment and north wing wall (left,
0°36'37.2''S, 80°25'29.5''W; right, 0°36'37.3''S, 80°25'29.5''W).

WALL IN BAHIA ALONG AVENIDA VIRGILIO RATTI

This wall in Bahía along Avenida Virgilio Ratti (0⁰36'30.6''S, 80⁰25'25.03''W) was
constructed using unreinforced stone and concrete masonry to contain the roadway prism and
protect from wave action in the bay. The exposed wall is ~1.5 m tall, with a thickness of 0.6 m
at the base and 0.3 m at the top. There is no information on the geometry or depth of the bottom
of the wall. The total wall height appears to have been raised by 0.5 m to facilitate raising the
grade and adding a sidewalk, which was done by building a second wall immediately behind
the original one. The retained soil (Fig. 7.5.15) appears to be silty clayey sand.

Approximately a 40-m long section of the wall collapsed as a result of the main April 16th
event, as shown on Fig. 7.5.15. The top of the wall appears to have rotated out by -10o
(counterclockwise) as measured at the section adjacent to the collapsed area (Fig. 7.5.16). The
seismic load and the negative batter of the wall increased the bending forces at the base, causing
the top of the wall to break off the foundation and topple. The wall sections adjacent to the
collapsed portion rotated, and the soil developed a shear plane about 1 m from the back of the
wall. The fill behind the wall settled approximately 0.5 m, which paired with a -10o rotation of
the wall. At the collapsed wall section, a secondary failure plane developed approximately 2
m behind the original wall and has also settled approximately 0.15 m (Fig. 7.5.17).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-88
Figure 7.5.15. Bahía wall along Avenida Virgilio Ratti: Collapse of unreinforced stone masonry of
(GPS: 0°36'29.8''S, 80°25'24.9''W).

Figure 7.5.16. Bahía wall along Avenida Virgilio Ratti: Wall deformation (left, 0°36'30.9''S,
80°25'25.2''W with GEER members R. Luque and G. Lyvers); Rotation and stacked walls (right,
0°36'30.3''S, 80°25'25.0''W with GEER members K. Rollins and C. Wood).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-89
15 cm

Figure 7.5.17. Bahía wall at Ave. Virgilio Ratti: Vertical deflection (left, 0°36'33.1''S, 80°25'25.6''W);
Wall bottom (center, 0°36'30.6''S, 80°25'25.0''W); Wall top (0°36'31.0''S, 80⁰25'25.2''W).

PORT OF MANTA WALL

At the Manta Port, the quay wall (0°56'28.9''S, 80°43'26.7''W) was constructed as a CIP
concrete wall, approximately 3-m tall and 0.2-m thick at the top. The wall supports the parking
lot that stores new imported cars (Fig. 7.5.18).

Figure 7.5.18. Aerial view of the collapsed wall at the Manta Port (GPS: 0°56'24.6''S, 80°43'22.0''W).

The lot experienced extensive liquefaction as described in Section 7.4. The wall rotated out
to a -4o (counterclockwise) batter and experienced lateral displacement of approximately 0.4
m at the top with the corner portion collapsed, as exhibited on Figs. 7.5.19 to 7.5.21. A boat
ramp in front of the wall heaved in the center about 1 m and the ramp concrete slabs buckled,
potentially due to a combination of liquefaction and lateral wall displacement.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-90
-4o
batter

Buckled
Boat
ramp
slab
b kl d Missing
wall corner

Figure 7.5.19. Manta Port wall: Damage and partial collapse (GPS: 0°56'28.5''S, 80°43'26.0''W).

Missing
corner
of wall

Buckling of
boat ramp
slab

Figure 7.5.20. Manta Port wall: Buckling of concrete boat ramp and collapse of wall corner (GPS:
0°56'29.1''S, 80°43'26.7''W).

~ 40 cm

Figure 7.5.21. Manta Port wall: Lateral displacement of 40 cm (GPS: 0°56'30.6''S, 80°43'28.0''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-91
WALL AT THE VELBONI SUPERMARKET IN PORTOVIEJO

The retaining wall at the Velboni supermarket in Portoviejo (1°3'37.8''S, 80°27'29.3''W)


was constructed with unreinforced brick masonry. With a total height of ~2 m, the wall only
retains 1 m of silty clayey sand as it was constructed on top of a sloped embankment to support
the parking lot of the supermarket. Before and after the earthquake photos are shown on Fig.
7.5.22, indicating that the wall itself did not fail, but rather the embankment failed along two
primary circular slip surfaces (Fig. 7.5.23). The slope on the same bank and on the other side
of the bridge also failed similarly.

Figure 7.5.22. Velboni supermarket wall in Portoviejo: Before (top, by Google Earth©) and after the
earthquake (bottom, by GEER team; GPS: 1°3'36.6''S, 80°27'29.6''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-92
2m

Figure 7.5.23. Wall at Portoviejo Velboni supermarket: Slope failure showing two primary slip surfaces
and wall rotating, following the slope movement (GPS: 1°3'37.8''S, 80°27'29.3''W). Team members A.
Caicedo (foreground) and A. Zekkos, K. Rollis, C. Wood, and an Ecuadorian colleague (background).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-93
7.6 Landslides and Rock Falls
INTRODUCTION

Hundreds of landslides and rock falls were recorded by the Ecuadorian Emergency
Operations Committee COE-3 after the main earthquake (Fig. 7.6.1). The observed slides
ranged from isolated rock falls to massive slope failures. The GEER team evaluated some of
these slides along the roads between Manta and Pedernales.

Figure 7.6.1. Landslides or rock falls reported by COE-3 along inspected roads.

OBSERVED LANDSLIDES AND ROCK FALLS

The reconnaissance team route is shown in Fig. 7.6.2, encountering a variety of landslides
and rock falls along the way. Generally, roads along the Ecuadorian coast are not designed
with any type of slope reinforcement. In some cases, a protection wall is built to safeguard the
roads, and potential slope instability is treated by flattening the slope.

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-94
Figure 7.6.2. Route followed during reconnaissance in blue line (Google Maps, ©2016).

OBSERVED FAILURES

Figures 7.6.3 to 7.6 depict failures encountered along the trail that was covered, either by
COE-3 immediately after the earthquake, or later by the GEER-ATC team.

Figure 7.6.3. Aerial photo of landslide on San Vicente-Canoa local road (GPS: 0°33' 5"S, 80°25'42"W).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-95
Figure 7.6.4. Landslide near Canoa. Photo by COE-3, dated 4/20/16 (0°30'15"S, 80°26'44"W).

Figure 7.6.5. Landslide near Canoa. (COE-3 photo, GPS: 0°30'15"S, 80°26'4"W).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-96
Figure 7.6.6. Landslide partially blocking the road near Jama (COE-3, GPS: 0°25'44"S, 80°26'59"W).

LANDSLIDES IN SAN ISIDRO, MANABÍ

Background

The annual rainy season in Ecuador typically extends from December to May and this year
was no different, with the El Niño phenomenon being added to the equation. The Ecuadorian
National Committee for Regional Study of El Niño (ERFEN) estimated an increment in the
precipitations intensity between end of March and beginning of April (Andes Newspaper,
2016) which were confirmed by abundant precipitations in the days before the earthquake. In
fact, the Ecuador Risk Management Secreteriat (SGR) estimated that flooding in coastal
regions due to heavy rains affected over 25,000 people, destroyed 140 houses, and damaged
over 4,000 more homes between January and the first two weeks of April 2016 (ECHO, 2016).
Based on these facts, SGR declared an orange alert for 19 of its 24 provinces for increased
rainfall (OCHA, 2016). A few cases of flooding and landslides were reported early in the year
(The Watchers, 2016), but it was not until mid-April that landslides became of bigger concern.

After the earthquake struck, coastal areas, particularly Manabí, were widely affected by
landslides. The British Geological Survey (BGS) was able to identify over 500 landslides in
the areas of Portoviejo, Bahía de Caráquez, Chone, Muisne and Crucita by comparing post-

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-97
earthquake satellite images (Pleiades) with pre-earthquake images from 2013-2015 Google
Earth (BGS, 2016). An estimated 28% of the identified landslides had a significant impact on
Ecuador’s infrastructure, as shown on the BGS pie chart of Fig. 7.6.7.

Figure. 7.6.7. Percentile of relative impact of earthquake-induced (“co-seismic”) landslides for a total
of 540 landslides observed in Portoviejo, Bahía de Caráquez, Chone, Muisne and Crucita (BGS, 2016).

Landslides Observations

Several landslides occurred in the province of Manabí. The landslides did not coincide with
known geologic faults in the area, and showed no evidence of the typical Angostura formation
sands at the Navas property. Instead, mudstone predominantly appeared at the surface after the
landslide succession. The geology map of Bahia de Caraquez (Fig. 7.6.8, Baldock, 1982) shows
geological formations, existing faults and approximate location of two macro-landslides of
interest. The two failures in land owned by Mr. G. Loor and Dr. C. Navas, were likely triggered
by increment in pore water pressure, soil strength reduction and/or liquefaction.

Figure 7.6.8. Location of Loor and Navas macro-landslides in geologic map (Baldock, 1982).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-98
Loor Site in San Isidro, Manabi

Approximately 290,000 m2 (29 ha) of the Loor site suffered extensive landslides and
ground deformations as shown on Fig. 7.6.9. Local residents observed water stored in tanks
was moving minutes before the main event. Consistently, a foreshock of Mw5.7 took place 11
min before the main event roughly at the same location as reported by IG-EPN (Fig. 7.6.10).

Aerial View Before 2016 Earthquake

Aerial View After 2016 Earthquake

Figure 7.6.9. Macro-landslides at Loor site (Ripalda, 2016) Base image: Google Earth ©2016.

The ground water was unusually high throughout this site, and could even be seen at the
bases of the slope due to the intense precipitation of the previous days (Fig. 7.6.11). The
landslide happened in a succession of three main events as illustrated in Fig. 7.6.12.

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-99
Figure 7.6.10. Events of 4/16/16: Mw5.7 foreshock (left) and Mw7.8 main (right) (IG-EPN, 2016).

Figure 7.6.11. Loor site: Ground water observed at slope base (GPS: 0°24'55.2"S, 80°14'50.3"W).

Figure 7.6.12. Sequence of landslides at Loor site (Ripalda, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-100
The first landslide was triggered by the main event and affected the lower and possibly the
middle zones of the slope. This landslide seemed to be translational at the top and rotational at
the base. The translational portion is evident due to lateral movement of soils, with trees
standing without losing their verticality. Rotated and fallen trees were identified in the
rotational portion, along with earth lifting at the base and stopping at the river. This landslide
left the upper portion, where the house and barn were located, without support, leading possibly
to a second translational landslide just few seconds after the first with ground uplift, rotation
and cracks around and through the house and barn. A third landslide took place with a rotational
mechanism as shown on Fig. 7.6.12. Rotation and settlement of the road and cracks around the
house, as well as bulging above the house near the head of the slope, were caused by this slide.

Combining many reconnaissance drone images and Google Earth photos prior to the
earthquake, our industry partner’s WSP | Parsons Brinckerhoff (WSP | PB) Virtual Design and
Construction team created a post-earthquake 3D model. A summary of the methodology used
(by T. Vaxevanis & J. Mezher of WSP|PB), data, and results are presented on Figs. 7.6.13 to
7.6.18. Figures 7.6.19 through 7.6.30 show damage from the landslide sequence.
DRONE IMAGES

DRONE IMAGES UPLOADED AND PROCESSED INTO SOFTWARE

3D SURFACE, 3D MODEL AND POINT CLOUD FILES ARE PRODUCED

DATA READY TO BE OVERPLAYED AND ANALYZED


Figure 7.6.13. Modelling methodology to create a 3D model from images taken by drones during
reconnaissance developed by the GEER-ATC (developed by WSP|PB, Virtual Design & Construction).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-101
Figure 7.6.14. Loor site prior to the earthquake used to develop the 3D model (Google Earth, ©2016).

Figure 7.6.15. Drone images of the Loor site landslides taken during reconnaissance after the
earthquake, used to develop the 3D model (Google Earth, ©2016).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-102
Figure 7.6.16. 3D model of Loor site developed by the GEER-ATC reconnaissance team digital photos and drone images (WSP|PB, Virtual Design & Construction).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-103
Figure 7.6.17. Section perpendicular hill slope at Loor site derived from the 3D model of Fig. 7.6.14. Approximate pre-earthquake slope included in dashed yellow line.

Figure 7.6.18. Section along hill slope at Loor site derived from the 3D model of Fig. 7.6.14. Approximate pre-earthquake slope included in dashed yellow line.

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-104
Figure 7.6.19. Loor site landslides: House at eastern end of the site destroyed in sunken area; secondary
(cane) house in lifted zone. Residual of white escarpment shale (GPS: 0°25'0.6"S, 80°15'40.1"W).

Figure 7.6.20. Loor site landslides: Destroyed ranch house and ground fissures around it (GPS:
0°25'0.1"S, 80°15'39.5"W).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-105
Figure 7.6.21. House sinking along with landslide at Loor site (GPS: 0°25'0.2"S, 80°15'39.1"W).

Figure 7.6.22. Loor site landslides: Evidence of ground uplift (GPS: 0°25'0.18"S, 80°15'39.53"W).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-106
Figure 7.6.23. Escarpment in the southwest side of Loor landslide (GPS: 0°25'2.3"S, 80°15'41.2"W).

Figure 7.6.24. Loor stables: Cracks (top, 0°25'1.4"S, 80°15'42.6"W; debris (bot., 0°25'0.8"S,
80°15'42.5"W).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-107
Figure 7.6.25. Evidence of ground uplifting in Loor landslide (GPS: 0°24'51.24"S, 80°15'50.34"W).

Figure 7.6.26. Debris on the east side of the Loor landslide (GPS: 0°25'0.6"S, 80°15'37.9"W).

Figure 7.6.27. North view of the extreme east of the landslide (GPS: 0°25'0.1"S, 80°15'38.2"W).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-108
Navas Site in San Isidro, Manabí

The Navas site suffered minor landslides with approximately 8.4 ha (vs. 29 ha at Loor)
exhibiting displacements (Fig. 7.6.28). The area is uninhabited, but the owner attested to rains
prior to the earthquake. The landslides affected roads at the slope top (Figs. 7.6.29, 7.6.30).

Navas site

Figure 7.6.28. Navas site macro-landslides (Ripalda, 2016). Base image: Google Earth (©2016).

Figure 7.6.29. Navas landslide west site with Isla Central at right. (GPS: 0°24'41.0"S, 80°13'28.1"W).

Figure 7.6.30. Photo at the eastern limit of the Navas landslide (GPS: 0°24'41.0"S, 80°13'28.0"W).

GEER-ATC Muisne, Ecuador, Earthquake 7- Geotechnical Observations


Report Version 1 7-109
7.7 Transportation Network
INTRODUCTION

The transportation network of Ecuador consists of highways, national roads, regional and
other roads (Fig. 7.7.1) of a total length of 43,670 km as of 2007 with 6,472 km paved and
37,198 are unpaved (CIA, 2016). The World Economic Forum ranked Ecuador as having one
of the best roadways in the Americas; second to the USA and at par with Canada (2015).

Figure 7.7.1. Ecuador Transportation Network (Academic, 2016).

The highway connecting the northern to the southern tip of Ecuador is called the “Ruta del
Sol.” Over the past eight years the Ruta del Sol and the rest of the roadway system have been
upgraded and reconstructed, resulting in significant improvement of the transportation network
quality for Ecuador. Since 2008 Ecuador’s rating by the World Economic Forum doubled from
2.6 to 5.2 (Fig. 7.7.2). These improvements were not only evident in the coastal roads but also
within the Sierra, thus improving the interconnectivity between towns and the coast. As part
of this improvement, Ecuador has also invested considerably in upgrading the existing bridges
as well as constructing new ones. Ecuador’s investment in its transportation network has been
approximately US$9 Billion since 2008 (Andes, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-110
6.0 5.7
5.2 5.2
4.9
5.0 4.6
4.2 4.3
3.9
4.0
3.6 3.6
3.2 3.3
3.1 3.1 3.0
3.0 2.8 2.9
2.7 2.7
2.3
2.0

1.0

0.0

Figure 7.7.2. Global Competitiveness Report (2015) Infrastructure: Quality of Road Ratings.

The GEER-ATC team traveled on 600 km of the nation’s roadway system through 200 km
from Guayaquil to Manta, 90 km from Manta to Portoviejo in both northern and southern
routes, 200 km from Manta to Pedernales, and 90 km from Manta to Chone. The team observed
cases of severe road damage, mostly near Portoviejo and by the coast between Bahia de Caracas
and Pedernales. The damage consisted primarily of settlement and lateral displacement of the
approach embankments as in the Mejia Bridge (Section 7.7), landslides such as the one in south
Canoa (Section 7.6), and concrete pavement cracking as on Fig. 7.7.3 south of Pedernales.

Figure 7.7.3. Concrete pavement damage south of Pedernales (GPS: 0°2'5.4"S, 80°4'27.3"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-111
The team observed extensive damage in one deadly bridge collapse in Guayaquil and two
bridges closed for repair that required significant construction intervention and long
interruptions, both for normal and emergency vehicles traffic and aid supplies. These were
isolated cases, since, of the 600 km covered, possibly 5-10 km seemed to have experienced
damage that required repair or replacement. While these cases generated interruptions in
operations, overall the majority of roads and bridges performed well and their serviceability
was maintained. In this section, observations on specific bridges are provided.

RIO CHICO (LAS CHACRAS) BRIDGE APPROACH

This bridge will be included in the next version of the report.

BRICENO BRIDGE APPROACH (GEOPIERS)

The Boca de Briceno Bridge is located next to the town of Boca de Briceno, 6 km south of
Canoa (Fig. 7.7.4). Figure 7.7.5 shows an aerial photo of the project.

Original U Shaped
Road Alignment

Figure 7.7.4. Briceno Bridge location. The original U-shape old roadway alignment was replaced by
this bridge and new alignment. (Google Earth Image ©2016 TerraMetrics).

Figure 7.7.5. Boca de Briceno Bridge aerial photo (GPS: 0⁰30'58.3''S, 80⁰26'29.1''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-112
The new roadway alignment was constructed from 2012 to 2014 to eliminate the U-shaped
section of the road and create a route without interfering with the scenic travel along the coast,
as part of a project extending from San Vicente to Pedernales that started in 2010. The new
alignment is about 1 km long and required construction of a new bridge to go over the Briceno
River and a 700-m approach embankment. The bridge is a 2-span structure with total length of
70 m, with the north span being 36.6 m and the south 33.4 m. The bridge is supported by a CIP
concrete cantilever abutment wall and a pier, both founded 10-m long drilled shafts with
diameter of 80 cm. Details are shown on the drawing of Fig. 7.7.6 (EMTOP, 2010b).

Figure 7.7.6. Boca de Briceno Bridge design layout (EMTOP, 2010b).

The embankment south of the bridge to the intersection is approximately 700-m long and
20-m wide at the top to accommodate 4 traffic lanes, sidewalk, and a bicycle path (Fig. 7.7.7).
The embankment height reaches 4.8 m with 2.5H:1V side slopes. A 60-cm thick construction
platform supported construction equipment and acted as permanent drainage layer.

An undated subsurface investigation by Asesorias y Estudio Tecnicos of 7 borings showed


a profile of loose silty sand to sandy silt from 2 to 5 m except at station 8+425, where a medium
stiff clay layer extends to 2.5 m depth, followed by stiff clay or sandy silt, dense silty or clayey
sand, sandy silt, and very stiff clay to 5 m. Below 5 m, stiff to very stiff clay was encountered.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-113
20 m

Figure 7.7.7. Boca de Briceno embankment geometry (GPS: 0°30'57.0"S, 80°26'30.1"W).

With the water table within 60 cm from the ground surface combined with presence of
loose silty sands, liquefaction was considered in the design. Geopiers, a ground improvement
technique, was implemented in the Boca de Briceno Bridge to address global stability and
settlement concerns, mitigate liquefaction and improve drainage for the embankment. Fig.
7.7.8 presents geotechnical parameters developed by Asesorias y Estudio Tecnicos.

Figure 7.7.8. Boca de Briceno Bridge geotechnical design parameters (Asesorias y Estudio Tecnicos).

Geopiers are compacted well-graded aggregate columns installed in pre-drilled holes (0.5
to 1 m in dia. to a max depth of 10 m) to form dense elements (Fig. 7.7.9). The aggregate is
pushed out in thin lifts laterally, densifying the surrounding foundation soil, with compaction
achieved using a high energy beveled tamper. This technology increases bearing capacity,
shear strength, rate of consolidation, liquefaction resistance, and reduces settlement. Geopiers
are applicable to different site conditions including soft organic clays, loose silts and sands,
uncompacted fill, stiff to very stiff clays, and medium dense to dense sands. Design parameters
include pier length, spacing, pier stiffness, and stress concentration ratio.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-114
Figure 7.7.9. Construction sequence using Geopiers. (Asesorias y Estudio Tecnico C. LTDA, undated).

A design Peak Ground Acceleration (PGA) of 0.4 g was considered as per contemporary
local standards for a hazard with an approximate return period Tr of 2,500 years or a probability
2% of being exceeded in 50 years. Applying this PGA to the proposed geometry and the soft
soil conditions resulted in global seismic factors of safety (FS) less than 1.1, which is the
minimum required by the local code. To achieve a FS ≥ 1.1, the slope would have to be
flattened to 3H:1V, even with the incorporation of the ground improvement. Hence, to optimize
cost by reducing the volume of fill and geopiers, an alternate 2.5H:1V slope scheme was
proposed that would provide a global stability FS = 1.08 for a PGA = 0.4 g. To meet the
required FS of 1.1, the PGA would have to be reduced to 0.37 g, which corresponds to a seismic
hazard of approximate T r = 2,100 years instead of the code-based 2,500 years. The designer
proposed this alternative that was accepted by the authorities on this project.

There was no strong ground motion instrument at the Boca de Briceno Bridge during the
main earthquake event. Based on the recordings from nearby strong ground motions shown on
Fig. 7.7.10, and considering the site conditions, the estimated PGA would be between 0.3 to
0.4 g, which is within the 0.37 g design PGA. Chapter 5 provides details on the recordings.

The liquefaction analysis in the Asesorias y Estudio report concluded a potential for the
silty sands to liquefy to a depth of 5 m at station 7+700 and to 3.5 m elsewhere. The impedance
contrast of the loose liquefiable soils with the compacted geopiers is large, transferring 4-45

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-115
times of the seismic load to the geopiers than in the soil. A transfer ratio of 12 was used in this
project, which significantly reduces the stress on the surrounding soil during a seismic event.

56 km

Figure 7.7.10. Boca de Briceno Bridge location (GPS 0°30'57.0"S, 80°26'30.1"W) with respect to
strong ground motion records the main earthquake epicenter. Base Image: Google Earth (©2016).

The settlement calculation utilizes conventional elastic theory using two-zone: an upper
and lower. The upper zone, is a combination of geopiers and soil and uses a composite modulus
to represent the geopier-soil system. For this calculation, the geopier and surrounding soil
moduli are needed, as well as the replacement ratio of the geopiers. In the lower zone the same
elastic theory is used to calculate settlements using the in-situ soil modulus. The total
settlement is the sum of the settlements calculated for each of the two zones.

The geopiers are constructed of a clean aggregate which reduces the drainage path distance,
hence increasing the rate of consolidation that can be accounted for in the settlement
evaluation. For this project the calculated settlements were estimated to be less than 2.5 cm 45
days after construction of the embankment.

The final embankment design based on the design criteria, site conditions, geometry, and
loading is shown on Figs. 7.7.11 and 7.7.12 with typical cross section and summary of column
spacing and depth based on location along the embankment (EMTOP, 2014). As-built
drawings reflect changes applied during construction; e.g., station 7+800 had some differences
from the original design. The as-built stone columns were 60 cm in diameter with variable
depths and spacing. Figure 7.7.13 shows photos of the geopier column installation and
equipment during construction.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-116
As discussed earlier, it is likely that the site experienced accelerations within design levels.
Liquefaction was evident near the embankment as shown on Fig. 7.7.14 in sand boils and sand
ejecta through cracks in the ground adjacent to the bridge abutment, with settlement around the
abutment footing shown on the bottom right photo.

Figure 7.7.11. Boca de Briceno Bridge: Geopier design layout by Station (Asesorias y Estudio Tecnico,
undated). Note English translation Espaciamiento (Spacing) and Profundidad (Depth).

Figure 7.7.12. Geopier design layout: Top - Typical section; Bot. - as built plans (EMTOP, 2014).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-117
Figure 7.7.13. Boca de Briceno Bridge construction: Left - equipment installing geopiers; right - top
of a completed 60-cm dia. geopier (est. GPS: 0°30'59"S, 80°26'29"W). Photos courtesy of Pivaltec.

Figure 7.7.14. Liquefaction and its effects next to Boca de Briceno Bridge embankment (top: est. GPS:
0°30'6"S, 80°26'3"W; bot. left: 0°30'59.1"S, 80°26'29.7"W; bot. right: 0°30'59.0"S, 80°26'28.9"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-118
The Boca de Briceno Bridge case history summarizes the satisfactory performance of its
embankment during a near-design level of PGA, with nearby evidence of liquefaction and
associated damage. Despite damage to neighboring infrastructure, this embankment exhibited
minimal damage while maintaining the serviceability of the road to the public. This can be
partially attributed to the ground improvement that met its objective to mitigate the liquefaction
effects immediately under the embankment by allowing pore pressures to dissipate, and
providing restrain to lateral spreading and settlement.

The GEER-ATC team visited the site on April 29, 2016 and observed minor damage to the
bridge, abutments, wing walls, as well as the embankment constructed on improved ground
(Fig. 7.7.15). Observations on the retaining wall are provided in Section 7.5. Along the
pavement, a (repairable) longitudinal crack about 5-15 cm wide, with 1-3 cm of vertical
displacement developed. Figure 7.7.16 shows no evidence of global vertical or lateral
movement in either side of the embankment, indicating that displacement was under the center.

Figure 7.7.15. Boca de Briceno Bridge: minor pavement damage (est. GPS: top 0°31'17"S,
80°26'24"W; bot: 0°31'12"S, 80°26'25"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-119
Figure 7.7.16. Boca de Briceno Bridge: overall good performance of the embankment with little to no
vertical displacement (GPS: 0°30'57.0"S, 80°26'30.1"W).

MANTA OVERPASS SITE

The Manta overpass site experienced strong ground shaking and surface evidence of
liquefaction was present at several locations (see Figs. 7.7.17 to 7.7.22). There was also surface
cracking of the pavement and sidewalk near a Mobil gas station and ground settlement around
the underground gas tanks. Geotechnical investigations performed by Geoestudios near the
Mobil gas station 6 days after the earthquake showed a loose silty sand layer with low plasticity
fines at depths of 1.5 and 3 m with SPT N-values varying from 2-5 bl/30 cm. The water table
was at 2 m, which may have caused for portion of the loose silty sand layer to liquefy during
the earthquake, resulting in the observed cracking and ground deformations.

Figure 7.7.17. Manta Overpass: Cracking and ground movement (GPS: 0 o57'1.2''S, 80o43'13.2''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-120
Figure 7.7.18. Cracked sidewalk at the Manta Overpass site (GPS: 0 o57'01.16''S, 80o43'13.24''W).

Figure 7.7.19. Possible evidence of sand boils and cracking due to lateral movement near the Manta
Overpass site (GPS: 0o57'01.7''S, 80o43'08.7''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-121
Figure 7.7.20. Cracking and ground movement at and around the Mobil gas station near the Manta
Overpass site (GPS: 0o56'59.1''S, 80o43'14.8''W).

Figure 7.7.21. Cracking and ground settlement around the underground tanks at the Mobil gas station
near the Manta Overpass site (GPS: 0o56'59.1''S, 80o43'14.8''W).

Figure 7.7.22. Cracking, ground movement north of Mobil station (GPS: 0o56'53.6''S, 80o43'16.0''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-122
MEJIA BRIDGE APPROACH

Introduction

The Mejia Bridge is composed of two bridges alongside, constructed roughly 15 years
apart, on the Portoviejo-Crucita road (Fig. 7.7.23), spanning ~50 m over the Portoviejo River
(Fig. 7.7.24). The bridges have a combined width of approximately 25 m, with 4 traffic lanes
in total. Based on design drawings, the abutments are supported by 25-m long driven piles and
the 20-cm concrete bridge deck is supported on steel beams (Figs. 7.7.25 to 7.7.27).

Figure 7.7.23. Location of Puente Mejia Bridge (Google Maps, ©2016).

Figure 7.7.24. Mejia Bridge prior to the earthquake (Google Maps, ©2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-123
Figure 7.7.25. Plan of the Mejia Bridges (EMTOP, 2011).

Figure 7.7.26. Typical longitudinal section of the Mejia Bridges (EMTOP, 2011).

Figure 7.7.27. Steel Beam and Concrete Deck Detail (EMTOP, 2011).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-124
Site Characterization

There are three subsurface investigations performed for this site. The first dates back to the
construction of the first bridge (“old bridge”), circa 2000, and consists of 54- and 60-m deep
borings at each abutment. The second investigation was done during construction of the second
bridge with a geotechnical report. The third investigation was conducted post-earthquake by
our Ecuadorian industry partners Geoestudios and Subterra, and consists of geophysical
MASW+ReMi, micro-tremors with SPT and CPTu at the locations are shown on Fig. 7.7.28.

Figure 7.7.28. Subsurface investigations at Mejia Bridge (Google Maps, ©2016).

The first investigation shows presence of different layers overlaying a strong, very dense
gravel layer which is a desiccated crust, with shear strength values between 60 and 180 kPa.
This layer extends to the water table, between 2.9 and 5.4 m below the ground level, and a high
plasticity soft to medium stiff clay and silt layer extends to approximately 20 m with variation
in shear strengths between 10 and 80 kPa. In between this layer, there is a 3-m layer of sandy
clay with an SPT N-value of 15 bl/30cm. A stiff clay and silt layer with shear strength from 80
to 200 kPa is present between 20 and 24 m, underlain by a clay layer with shear strength
between 60 and 200 kPa, estimated from SPT N-values. The plasticity index (PI) ranges

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-125
approximately from 25 to 52. A summary of the boring results are shown below on Figs. 7.7.29
and 7.7.32.

Figure 7.7.29. Pre-earthquake Boring B-1 at Mejia Bridge site.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-126
Figure 7.7.30. Pre-earthquake Boring B-2 at Mejia Bridge site.

Modifications and improvement work was performed to the road in 2007. According to the
project drawings, Mejia Bridge is located at Station 6+602.4 within Zone 1 of the project which
extends from Sta. 0+000 to Sta. 7+750 (Technical University of Manabi). Based on the
geotechnical report from the 2 nd investigation, the silty clays and clayey silts have PI between
11 and 28, increased near the river from 32 to 36. The soil conditions are fairly consistent to
the first investigation that seems to have underestimated the plasticity of the clay layers.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-127
The last set of data was obtained from the 3rd investigation following the main earthquake
of April 16th approximately 50 m south of the bridge, towards Portoviejo. One boring (P-01)
with SPT was drilled from the top of the embankment on the east side and one CPTu was
performed on the side of the embankment, on the west side of the road. The boring log (Fig.
7.7.31) showed subsurface of silty sand with some layers of sandy silt with SPT N-values 15
to 31 bl/30cm over the first 4 m, down to values between 10 and 17 bl/30cm at depths of 4 to
12 m. From 12 to 19 m the N-values were generally greater than 40 bl/30cm.

Figure 7.7.31. Post-earthquake Boring P-01 at Mejia Bridge site.

The CPTu classified the soil mostly as a clay and silty clay with some sand layers between
10 and 14 m of depth. The shear strength varied from 20-50 kPa at the top 5 m, increased up
to 150 kPa between 14 and 17 m depths (Fig. 7.7.32). This variation is consistent with the
findings of the first borings where the shear strength was higher than the average of the profile
at depths from 20-24 m. The CPTu coordinates place it not on the road, but rather the side of
the embankment, approximately 4 m below the road level. Based on the ground water levels,
the top of the boring and the CPTu is at different elevations with respect to the “old borings,”
with the CPTu about 1.5 to 4 m lower. Results of Vs measurements are shown on Fig. 7.7.33
from the CPTu and on Fig. 7.7.34 from ReMi+MASW with Vs,30 of 176 and 214 m/s.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-128
Figure 7.7.32. Mejia Bridge site: Post-earthquake CPTu results.

Figure 7.7.33. Mejia Bridge site: Post-earthquake Vs measurements from CPTu.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-129
S-wave velocity
S-wave (m/s)
velocity (m/s)

0 020 20
40 40
60 60
80 80
100100
120120
140140
160160
180180
200200
220220
240240
260260
280280
300300
320320
340340
0 0 S-wave velocity (m/s)
Shear Wave Velocity Vs (m/s)
2 2
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280 300 320 340
4 4
0
6 6
2
8 8
4
10 10
6
12 12
8
Depth (m)
Depth (m)

14 14
10
16 16
12
18 18
Depth (m)

14
20 20
16
22 22
18 S-wave velocity (m/s)
24 24 S-wave velocity (m/s)
20
0 S-wave
20 velocity
40
S-wave 60 model
velocity (inverted)
80model
100 120 140: REMI
(inverted) :160 -180
REMI MASW
2001.rst
- MASW220 240 260 280 300
1.rst
220 0 20Vs40 60= 176.5
80 100 120 140 160 180 200 220 240 260 280 300 320 340
Average
Average 30m
Vs 30m m/sec
= 176.5 m/sec
242 0 S-wave velocity (m/s)
2S-wave velocity model (inverted) : REMI - MASW 1.rst
4 0 20 40 60 80 100 120 140 160 180 200 220 240 260 280 300
0 4Average Vs 30m = 176.5 m/sec
6
2 6
8
4 8
10
6 10
12
8 12
Depth (m)

14
Depth (m)

10 14
16
12 16
18
14 18
Depth (m)

20
16 20
22
18 22
24
20 24
S-wave velocity
S-wave model
velocity (inverted)
model : REMI-MASW
(inverted) 2.rst
: REMI - MASW 1.rst
22
Average Vs Vs
Average 30m30m= 214.1 m/sec
= 176.5 m/sec
24
S-wave velocity model (inverted) : REMI-MASW 2.rst
Figure 7.7.34. Mejia Bridge site: Post-earthquake Vs measurements from ReMi+MASW.
Average Vs 30m = 214.1 m/sec

Approach Embankment Failure

After the main event, the embankment slope of the Mejia Bridge approach failed, causing
the road to collapse as shown on Figs. 7.7.35 and 7.7.36. The failure seems to have diverted
part of the river which showed signs of erosion. The gabion also failed, likely not designed for
the additional forces from the soil deformation.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-130
Based on the reconnaissance data, it appears that on the side where the embankment failed,
the soils are predominantly clays and silty clays. However, the investigations indicate a
saturated surficial silty clay layer (old borings) and of sandy silt (new borings) with SPT N-
values between 7 and 10 bl/30cm. Excess pore pressures during the earthquake could have
developed in this layer and caused the road movement.

Figure 7.7.35. Mejia Bridge approach embankment: Aerial views of failure by COE-3 (GPS: top,
0°59'24"S, 80°28'11"W; bottom, 0°59'24"S, 80°28'11"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-131
Figure 7.7.37. Mejia Bridge: Road displacement (GPS: 0°59'24"S, 80°28'11"W, Geoestudios).

THE BAHIA-SAN VICENTE, LOS CARAS BRIDGE

Figure 7.7.38. Bahia-San Vicente, Los Caras Bridge. Photo by Ecuador Army Corps of Engineers
looking eastward from San Vicente to Bahía de Caráquez (GPS: 0°36'30.7"S, 80°24'30.6"W).
GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations
Report Version 1 7-132
The Los Caras (Bahía) Bridge, shown on Fig. 7.7.38 is a critical transportation connection
crossing the River Chone Estuary and joins the coastal towns of San Vicente and Bahía de
Caráquez, in the Manabí province. The project was undertaken by Ecuador Army Corps of
Engineers and was built in 2010. The bridge consists of a central span from Abutment E-1 to
Abutment E-2 for a total length of 1,984.5 m. The entry to the bridge is via a 208-m access
ramp from the Bahía span, and a 607-m access ramp from the San Vicente span. The bridge is
supported by 48 piers and two abutments divided in three sections shown on Fig. 7.7.39:

 Bahía de Caráquez Access: Abutment E-1 and Piers P-1 through P-6, a total length L of
120.3 m and width W of 10.2 m.
 Central Section: Piers P-6 to P-44, total L = 1,710 m with span of 45 m between the 38
piers. There are 2 types of continuous lengths, 4 continuous spans, with L = 180 m, and 2
continuous spans, with L = 90 m and W = 13.2 m.
 San Vicente Access: Piers P-44 to P-48, Abutment E-2 with L = 154 m and W = 13.2 m.

Figure 7.7.38. Aerial view of Los Caras (Bahía) Bridge by Ecuador Army Corps of Engineers (GPS:
0°36'33.6"S, 80°24'58.7"W).

The bridge superstructure is designed with pre-stressed pre-tensioned and post-tensioned


concrete. The bridge deck is 9.3 m wide for vehicular traffic and has a 3 -m wide pedestrian
and bicycle lane. The roadway section has 2 traffic lanes. Each of the bridge’s concrete slab
spans consists of 6 beams, and of these, 2 are directly supported on seismic isolators (152 Triple
Pendulum bearings in the central section). The bridge cross section is shown on Fig. 7.7.39.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-133
Figure 7.7.39. Los Caras (Bahía) Bridge cross-section (Ecuador Army Corps of Engineers, 2010).

The site is characterized as an estuarine environment formed by alluvial deposits with


depths between 25 and 85 m overlying weathered rocks (Fig. 7.7.40 by EC Army Corps of
Engineers and A. Caicedo). The foundations are open-end steel friction pipe piles with
diameter of 1.21 m, wall thickness of 20 mm and length varying from 32 to 65 m.

Figure 7.7.40. Los Caras (Bahía) Bridge soil profile (Ecuador Army Corps of Engineers, 2010).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-134
The river piers were designed to withstand earthquake loading through use of triple friction
pendulum seismic isolation devices that exhibit significant non-linear behavior (Constantinou
et al., 2011). The devices are designed to prevent transfer of large seismic forces to the
superstructure by decoupling it from the substructure (Christopoulos et al., 2006). The overall
result is a reduction of inertial forces in the superstructure and substructure, and redistribution
of forces between the piers and the abutments (Constantinou et al., 2007). The isolator’s
maximum lateral displacement capacity is 58.4 cm according to the design by the Ecuador
Army Corps of Engineers. Conventional elastomeric bearings were used in the approaches.

Post-Earthquake Observations

Observers stated that the bridge slithered “like a snake” during the main earthquake. Based
on visual inspection comments by engineer Marcelo Romo who was responsible for the design
as member of the Ecuador Army Corps of Engineers (ECACE), the bridge was operable with
no evidence of damage that would require closure. In Fig. 7.7.41, members of the ECACE
show the inspection by the Ministry of Urban Development and Housing (MIDUVI).

Figure 7.7.41. Ecuador Army Corps of Engineers team members show the MIDUVI inspection
document of the Los Caras (Bahía) Bridge (GPS: 0°36'48.9"S, 80°24'46.9"W).

There was no visible damage to the modular long expansion joint at the west (Bahía de
Caráquez) end of the bridge (Fig. 7.7.42). At the east (San Vicente) end, there was a detachment
of a neoprene section and expansion of the joints (Fig. 7.7.43). Near the middle of the bridge,
there was no evidence of damage (Fig. 7.7.44). There was no significant cracking of the
concrete deck which might have indicated settlements of the piers.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-135
Figure 7.7.42. Los Caras (Bahía) Bridge: Modular expansion joint, Bahía de Caráquez side (GPS:
0°36'48.85"S, 80°24'46.93"W), by Romo (2016).

Figure 7.7.43. Los Caras (Bahía) Bridge: Modular expansion joint, San Vincente side (GPS:
0°36'30.74"S, 80°24'30.54"W), by Romo (2016).

Figure 7.7.44. Los Caras (Bahía) Bridge: Neoprene rubber standard expansion joint (GPS:
0°36'35.28"S, 80°25'17.15"W), by Romo (2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-136
Figure 7.7.45 illustrates the impact absorbers (external movement synchronizers) that
reflected at least 5 major impacts in the earthquake. The elastomeric bearings of the Bahía and
San Vicente approaches did not show evidence of damage (Romo, 2016). There were cracks
on the horizontal shear blocks in Piers 3, 4 and 5 (south end of Bahía de Caráquez access) that
needed repairs, but were able to restrain lateral movement of the superstructure (Fig. 7.7.46).

Figure 7.7.45. Los Caras (Bahía) Bridge: External movements synchronizers (GPS: 0°36'33.6"S,
80°24'58.7"W), by Romo (2016).

Figure 7.7.46. Los Caras (Bahía) Bridge: Cracking of Pier 3 on the horizontal shear block (Bahia side).
The pier has elastomeric bearings (GPS: 0°36'36.6"S, 80°25'27.3"W), by EC Army Corps of Engineers.

All piers had relatively uniform behavior in the earthquake, except for Pier 12 and its
neighbors. The bearings indicate that lateral displacement took place. The bearings towards the
side face of the bridge appear to show a larger lateral deformation by about 50% from south to

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-137
north. Pier 12 was the only pier that showed much larger displacements (Romo, 2016). The
base isolation system did show damage, according to the inspection carried out by the team of
the Ecuadorian Army Corps of Engineers. The average differences were in the range of 15 cm
in the north direction and 30 cm in the southern direction (Fig. 7.7.47).

Figure 7.7.47. Friction pendulum device as part of the base isolation system of Los Caras Bridge (Left),
Inner displacement isolator view (GPS: 0°36'33.58"S, 80°24'58.68"W), by Romo, 2016.

AVENIDA DE LAS AMERICAS OVERPASS, GUAY

This section will be included in the next version of the report.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-138
7.8 Dams
INTRODUCTION

After the main seismic Mw7.8 event of April 16 th, 2016, the Ministry of Urban
Development and Housing (MIDUVI) in coordination with the Public Water Utility (EPA-EP),
organized inspections of four dams (La Esperanza, Poza Honda, Chone, and San Vicente) in
the Manabí province and the northern part of Santa Elena. These inspections were not only
necessary due to the earthquake, but also due to heavy rains in the area on April 10 th and 11th.

The inspections were performed on April 26 th, 2016 by a group of geotechnical and
hydraulic engineering experts. The commission was composed of the following professionals:
Carlos Villamarín, Ricardo Paredes and Joaquín Negrete, EPA officials and members of the
Commission on Dams of the Technical Committee No. 3 of MIDUVI, consultants Oswaldo
Ripalda and Pedro Castro, from the Catholic University of Santiago de Guayaquil (UCSG),
and consultant C. Blas Cruz. The locations of the four inspected dams are shown on Fig. 7.8.1.

Figure 7.8.1. Locations of four inspected dams as yellow pins. Base image: Google Earth (©2016).

Part of the GEER-ATC team, led by Gabriela Lyvers of the US Army Corps of Engineers,
visited the Chone Dam construction site on April 30 th, 2016. The group included K. Rollins
(Brigham Young Univ.), C. Wood (Univ. of Arkansas), D. Alzamora (Federal Highway
Administration), and A. Caicedo (UCSG). Results from both inspections are presented in this
section, where all left and right designations are made looking in the downstream direction.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-139
LA ESPERANZA DAM

History and Design Information


La Esperanza Dam is a multipurpose dam completed in September 1995, located on the
Carrizal River in the Province of Manabí, approximately 150 km from the city of Guayaquil
with geographic coordinates (0° 53' 18.3" S, 80° 4' 34.4" W). An aerial view is shown on Fig.
7.8.2. This facility serves the purposes of flood control, irrigation, and water supply for both
industrial and city use. The dam is owned and operated by the National Secretariat of Water
(Secretaría National de Agua), known as SENAGUA.

Figure 7.8.2. La Esperanza Dam aerial view from helicopter flyover (GPS: 0°53' 2.5''S, 80°5'3.0''W).

The dam is 47 m in height with crest length of 700 m, built as a zoned embankment with a
clay core and a sand and gravel shell. A cross section and plan view are provided on Figs. 7.8.3
and 7.8.4. Subsurface investigations performed in the early 1980’s indicated that the dam
foundation is comprised of shale with signs of ancient and more modern landslides, underlain
by a shear zone. Some of these landslides were thought to be active.

At the center of the dam, the old river channel was filled with more recent alluvial silty
sands interbedded with weak clays. The main concern during construction was differential
settlement between the dam founded on rock and its central section founded on the alluvium.
This concern drove the decision to redesign the embankment prior to construction in early
1990’s.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-140
Figure 7.8.3. La Esperanza Dam: Embankment cross section (Rehabilitation Center, Manabí, 1983).

Figure 7.8.4. La Esperanza Dam: Plan view (Rehabilitation Center, Manabí, 1983).

Drainage galleries were extended from the right and left abutment as seen on Fig. 7.8.5.
Drainage was extended beyond the shear zone into the shale above. The outlet works of the
dam are located near the right abutment and consist of an intake tower, conduit, and stilling
basin, shown on Fig. 7.8.6. Detailed drawings of the cross sections were not available.

A spillway at the right abutment is present and controlled by 4 radial gates as seen on Fig.
7.8.7. Based on available information, the spillway and conduit are both founded on shale.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-141
Figure 7.8.5. La Esperanza Dam: Abutment drainage galleries (Rehabilitation Center, Manabí, 1983).

Figure 7.8.6. La Esperanza Dam: Intake tower and conduit drawings (Rehab. Center, Manabí, 1983).

Figure 7.8.7. Spillway section at right abutment (Rehabilitation Center, Manabí, 1983).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-142
Dam Instrumentation
Accelerometer data near the site will be provided in the next version of this report. The
team is in search of any other available instrumentation data at the site.

Inspection Observations
The MIDUVI inspection commission reported no signs of distress of the embankment due
to the main seismic event. No cracks, deformations or settlements were observed at the crest
or slopes. No increase of seepage at the toe of the dam was reported either.

Dam personnel reported that there was an increase in seepage in the drainage gallery after
the main event. The MIDUVI team inspected both the right and left drainage galleries (Fig.
7.8.8). The team was only able to inspect approximately half the length of the right and left
drainage gallery due to overflowing pump wells. It was reported by project personnel that there
was no sign of structural distress of the galleries due to the seismic event. According to
drainage flow rates after the seismic event, the flow rate increased by 20% at the right gallery
and 60% at the left. The flow rates at the time of the inspection were 3.8 and 3.9 l/s for the
right and left galleries, respectively. Since there were heavy rains on April 10 and 11, 2016, it
is unknown if the seismic event further increased the seepage.

Figure 7.8.8. La Esperanza Dam: Right and left drainage galleries after the main seismic event. Photos
from the MIDUVI Inspection Commission (GPS: 0° 53' 19.0'' S, 80° 4' 35.3'' W).

Some damage was observed between the intake tower and the access bridge with
differential settlement estimated between 3-4 cm (Fig. 7.8.9). Damage to the intake tower with
some broken windows was observed. Power lines went down after the event, but were repaired
shortly after, with power at the intake tower restored at the time of the inspection. Columns
supporting the gate hoist were damaged (Fig. 7.8.10), although the gate remained operable.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-143
Figure 7.8.9. La Esperanza Dam: Differential movement of intake tower and access bridge. Photo from
the MIDUVI Inspection Commission (GPS: 0° 53' 14.3'' S, 80° 4' 30.4' W).

Figure 7.8.10. La Esperanza Dam: Damage to columns supporting the intake tower gate hoist. Photo
from the MIDUVI Inspection Commission (GPS: 0° 53' 14.5'' S, 80° 4' 30.0'' W).

The spillway at this dam is controlled by 4 radial gates. The spillway section is shown on
Fig. 7.8.11. At the time of the MIDUVI inspection, the radial gates were open approximately
0.5 m as shown on Fig. 7.8.12. It is unknown if the pool was loading the spillway gates during
the main seismic event. No issues at the spillway section were reported.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-144
Figure 7.8.11. La Esperanza Dam: Spillway radial gates. Photo from the MIDUVI Inspection
Commission (GPS: 0° 53' 9.2'' S, 80° 4' 23.6'' W).

Figure 7.8.1. La Esperanza Dam: A spillway radial gate. Photo from the MIDUVI Inspection
Commission (GPS: 0°53' 8.9'' S, 80° 4' 36.4'' W).

No issues were reported at the stilling basin by the MIDUVI Inspection Committee. The
conduit was not inspected following the main earthquake.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-145
POZA HONDA DAM

History and Design Information

Poza Honda Dam is located 2 km from Las Mercedes and 30 km from the city of Portoviejo,
in the Manabí province with geographic coordinates (1° 6' 49.5" S, 80° 12' 13.5" W). An aerial
view of this zoned embankment dam is shown on Fig. 7.8.13. The dam was completed in 1971
with height of 40 m and crest length of 270 m, with primary functions for irrigation and water
supply for the cities of Rocafuerte, Portoviejo, Sucre, Jipijapa, and Jaramijó.

Figure 7.8.13. Poza Honda Dam aerial view. Base Image: Google Earth (©2016).

The dam has an uncontrolled spillway on the left abutment (Fig. 7.8.14) with 70 m of total
downstream length and max capacity of 875 m³/s. The outlet works are at the right abutment
and the conduit, controlled by 2 shutoff valves, has diameter of 3 m and total length of 300 m.

Figure 7.8.14. Poza Honda Dam. MIDUVI Inspection photo (GPS: 1 °6' 39.8'' S, 80° 12' 32.5'' W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-146
Dam Instrumentation
Several Casagrande standpipe piezometers and survey monuments are available at the Poza
Honda Dam site. It is unknown if baseline data and readings from the days following the main
earthquake exist. The next version of this report will contain any such data if available and
accelerometer recordings near the site.

Inspection Observations
The MIDUVI Inspection Commission reported some visible settlement at the crest of the
dam next to the right spillway training wall. The differential settlement was measured to be 3
cm at the time of inspection. Small transverse cracks were seen at the embankment crest, but
they were found to be shallow as seen on Fig. 7.8.15. Dam personnel reported that these cracks
were not present before the seismic event. The MIDUVI team reported visible cracks in the
concrete slabs at the upstream face (Fig. 7.8.16) that were present before the earthquake
according to dam personnel. During inspection, no seepage was observed at the contact
between the embankment and the retaining wall. No other issues were reported at the crest.

Figure 7.8.15. Poza Honda Dam: Transverse cracks at crest. MIDUVI Inspection Commission photo
(GPS: 1° 06' 51.7'' S, 80° 12' 16.5'' W).

At the downstream face no issues were noted. No leaks, cracks or increased seepage was
seen at the toe or embankment/spillway interface. Figure 7.8.17 shows the drainage ditch at the
downstream face with no seepage at the inspection time. No additional seepage was observed
from the relief wells (see Fig. 7.8.18).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-147
The MIDUVI team also inspected the uncontrolled spillway, stilling basin and spillway
bridge and no issues were reported. The stilling basin was not dewatered at the time of
inspection as can be seen on Fig. 7.8.19.

Figure 7.8.16. Poza Honda Dam: Cracks in upstream concrete slabs. MIDUVI Inspection Commission
photo (GPS: 1° 6' 51.8'' S, 80° 12' 15.8'' W).

Figure 7.8.17. Poza Honda Dam: Drainage ditch at toe with no seepage at the time of inspection.
MIDUVI Inspection Commission photo (GPS: 1° 6' 51.1'' S, 80° 12' 17.1'' W).

During the time of the MIDUVI inspection, the low flow outlet was being used at maximum
capacity with an outflow of 30 m3/s to control reservoir pool elevation (see Fig. 7.8.20). No
further inspection or dewatering was done at this time.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-148
Figure 7.8.18. Poza Honda Dam: No seepage observed in the relief wells at the time of inspection,
MIDUVI Inspection Commission photo (GPS: 1° 6' 46.7'' S, 80° 12' 17.8'' W).

Figure 7.8.19. Poza Honda Dam: Uncontrolled spillway and stilling basin. MIDUVI Inspection
Commission photo (GPS: 1° 6' 52.2'' S, 80° 12' 17.0'' W).

Figure 7.8.20. Poza Honda Dam: Low flow outlet at time of inspection. MIDUVI Inspection
Commission photo (GPS: 1° 6' 46.2'' S, 80° 12' 17.8'' W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-149
CHONE/RIO GRANDE DAM

History and Design Information

Chone (Rio Grande) Dam is a multipurpose dam currently under construction since 2011
in the Manabí province, expected to be completed and transferred to the government in few
months. The dam is located downstream of the confluence of the Platanales and Chone Rivers,
approximately 15 km upstream from the town of Chone with geographic coordinates (0° 41'
59.6" S, 79° 59' 21.5" W). An aerial view of the dam from the GEER-ATC helicopter flyover
is shown on Fig. 7.8.21. The Chone Dam is owned and operated by SENAGUA with authorized
purposes of flood management, irrigation, and water supply for both industrial and city use.

Figure 7.8.21. Chone (Rio Grande) Dam aerial view from flyover (GPS: 0°41'44.1''S, 79°59'30.6''W).

The foundation of the Chone Dam is composed of horizontally bedded siltstones and
sandstones, as show on the geologic profile of Fig. 7.8.22. The Borbón Formation of medium-
grained sandstones interbedded with fine-grained siltstone is found from Elev. 179 to 90 MSL.
From elevation 90 to 30 MSL, there is a transition zone between the Borbón and underlying
Onzole formation composed of alternating layers of fine grained sandstone and siltstone.
Below Elev. 30 MSL, the Onzole Formation is an alternating siltstone/shale layer.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-150
Figure 7.8.22. Chone Dam: Geologic section after Chávez (2013).

This facility is a high risk dam as approximately 25,000 people live at a short distance
downstream including the towns of Chone and San Antonio. It is a zoned embankment dam
with maximum height of 57.5 m and capacity of 76 hm3 once construction is complete. The
crest is at Elev. 72.5 MSL with total crest length of 276 m and width of 8 m. The dam has a
195-m long uncontrolled spillway at the left abutment. The spillway chute is angled at 36 o and
flows into a 17-m stilling basin. The low flow outlet is through a tunnel which outflows into
the spillway stilling basin. The tunnel is controlled by 2 slide gates, each 1.5 1.5 m2. The
spillway is founded in the Borbón-Onzole transition, and the diversion tunnel and outlet will
be founded on the Onzole formation.

Figure 7.8.23. Chone Dam: Plan view after Chávez (2013).


GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations
Report Version 1 7-151
During the 2013 design re-evaluation, Dr. M.A. Chávez Moncayo noted that the subsurface
soil layers are almost horizontal within the valley where the dam will be located, which is
indication that despite some historic tectonic movement and presence of faults, these did not
cause significant deformations. Original design documentation is available by ACOLIT
(2008), with updated drawings on Figs. 7.8.23 and 7.8.24 from the 2013 study. Foundation
design includes a grout curtain extending from each abutment to the dam center, overlapping
a cutoff wall that runs 56 m from each abutment to the dam center as shown on Fig. 7.8.25.

The 2013 design re-evaluation determined that the dam would remain stable for PGA up
to 0.5 g. Although the cuts at the right and left abutment were deemed stable under hydrologic
and seismic loading, the stability was considered temporary based on the erodibility of the
exposed rock. The jointing in the rock has the ability to form moveable blocks and the stability
should be evaluated as conditions of the cut deteriorate over time.

Figure 7.8.24. Chone Dam: Typical dam cross section after Chávez (2013).

Figure 7.8.25. Chone Dam: Grout curtain and cutoff wall after Chávez (2013).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-152
Dam Instrumentation
The Chone Dam is instrumented with inclinometers, accelerometers, settlement plates,
survey monuments, vibrating wire piezometers and standpipe piezometers (Fig. 7.8.26).
Instrumentation and accelerometer data from the time of the main earthquake, if become
available through the construction company, will be included in the next version of this report.

Figure 7.8.26. Instrumentation at Chone Dam (ESPE-Innovativa & Empresa Publica del Agua, 2015).

Inspection Observations

Following the MIDUVI inspection of the Chone Dam, members of the GEER-ATC team
visited the facility and took in-situ shear wave velocity Vs measurements using surface wave
techniques and HVSR that are being evaluated and will be reported in the next version of this
report. The next paragraphs present observations from the MIDUVI and GEER-ATC teams.

At the dam embankment, no settlement or deformations were visible during inspection.


Leaks at the downstream face and spillway contact were not visible either. The dam is currently
not loaded by a reservoir since construction was just completed. Filling of the reservoir is
expected to occur during the next rainy season. Water was present at the toe of the dam at the
time of the inspection as seen on Fig. 7.8.27. With strong rains just few days before the
earthquake, it is likely that this wet area was due to the rain, but it is unknown if there were
any changes in seepage at the toe shortly after the seismic event.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-153
Figure 7.8.27. Chone Dam: Wet spot at the toe (GPS: 0° 41' 43.2'' S, 79° 59' 29.6'' W).

Figure 7.8.29. Chone Dam: Concrete spillway (GPS: 0° 41' 50.5'' S, 79° 59' 36.6'' W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-154
No significant damage was observed at the spillway, chute or stilling basin, that were not
in use at the inspection time (Fig. 7.8.28), but there was water at the stilling basin which
reduced visibility. Some cracks at the spillway training walls, counterfort and slab were visible
(Figs. 7.8.29, 7.8.30), but it is unknown when they occurred. Some cracks had vegetation
growth, so they may have pre-existed the earthquake, although this cannot be confirmed.

Figure 7.8.29. Chone Dam: Spillway cracks in slab, wall, counterfort (GPS 0°42'2.0''S, 79°59' 25.9''W).

Figure 7.8.30. Chone Dam: Spillway crack in wall and counterfort (GPS: 0° 42' 2.1''S, 79° 59' 25.6''W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-155
There was some damage visible to the gate seal of the low flow (Fig. 7.8.31), which
according to dam personnel, had occurred prior to the earthquake and a replacement was
planned. Grout injections were also planned in the 3 outlet tunnel before the event to mitigate
leakage. The outlet tunnel which is 3 m in diameter was not dewatered or inspected.

Figure 7.8.31. Chone Dam: Gate seal at low flow outlet (MIDUVI, GPS: 0° 42' 5.6'' S, 79° 59' 21.5''W).

Slope instability at the cuts at the right and left abutment were visible (Figs. 7.8.32 to
7.8.35). Based on the dam personnel, some instability had occurred due to the heavy rains
preceding the seismic event, which may have escalated with the earthquake.

Figure 7.8.32. Chone Dam: Slope instability at right abutment (GPS: 0° 41' 40.7'' S, 79° 59' 17.7' 'W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-156
Figure 7.8.33. Slope instability at right abutment above dam (GPS: 0° 42' 1.6'' S, 79° 59' 10.9'' W).

Figure 7.8.34. Slope instability, right abutment, above dam. MIDUVI Inspection Commission photo
(GPS: 0° 41' 59.0'' S, 79° 59' 21.1'' W).

Figure 7.8.35. Slope instability, left abutment by concrete spillway. MIDUVI Inspection Commission
photo (GPS: 0°42'5.0'' S, 79°59'25.6'' W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-157
SAN VICENTE DAM

History and Design Information

San Vicente Dam is located at the northern section of the Santa Elena province,
approximately 70 km from Guayaquil. The dam facilitates irrigation and water supply to the
coastal area. An aerial view and the spillway are shown on Fig. 7.8.36 and 7.8.37, respectively.

Figure 7.8.36. San Vicente Dam: Aerial view. Base Image: Google Earth (©2016).

Figure 7.8.37. San Vicente Dam: Uncontrolled spillway. Base Image: Google Earth (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-158
Dam Instrumentation
Accelerometer data, when available, will be added to next version of this report.

Inspection Observations
The inspection of this dam was only done by Oswaldo Ripalda on April 30 th. No cracks,
fissures, settlement or deformations were observed on the embankment. At the toe of the dam
or at the spillway/embankment interface no signs of failure were observed. No recent damage
to the spillway, chute or stilling basin was observed during inspection. This dam component
was not in operation during the seismic event since the reservoir pool level was lower than the
elevation of the spillway crest (Fig 7.8.38). Before the seismic event, the joints of the spillway
slabs had been repaired due to joint separation of approximately 10 cm at some locations as
seen in Figs. 7.8.39 and 7.8.40. No issues were found with the recently repaired joints.

Figure 7.8.38. San Vicente Dam: Uncontrolled spillway crest. MIDUVI Inspection Commission photo
(GPS: 2° 0' 1.8'' S, 80° 31' 54.8'' W).

The spillway training walls had been repaired in the past due to cracking at the joints that
appeared to be due to differential displacement of the wall sections and the location of the
concrete key at the wall joints. The cracks seem to have reopened after the main seismic event
(Figs. 7.8.41, 7.8.42). The spillway bridge was also inspected at the same (Fig. 7.8.43). Lateral
outward movement was observed at the left and right bridge supports. Separation was
approximately 20 cm at the left support as can be seen on Fig. 7.8.44. Separation at the right
support was measured to be 7 cm as shown on Fig. 7.8.45.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-159
Figure 7.8.39. San Vicente Dam: Spillway chute slabs (MIDUVI, GPS: 2° 0'’ 1.6'' S, 80° 31' 56.1'' W).

Figure 7.8.40. San Vicente Dam: Repaired spillway joint (MIDUVI, GPS: 2°0'1.6''S, 80°31' 56.1''W).

Figure 7.8.41. San Vicente Dam: Shear crack at spillway wall joint, (MIDUVI, GPS: 2° 0' 1.4'' S, 80°
31' 54.6'' W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-160
Figure 7.8.42. San Vicente Dam: Shear crack at spillway training wall joint. Top of MIDUVI Inspection
Commission photo (GPS: 2° 0' 1.3'' S, 80° 31' 55.8'' W).

Figure 7.8.43. San Vicente Bridge, Santa Elena. MIDUVI Inspection Commission photo (GPS: 2° 0'
1.8'' S, 80° 31' 54.4'' W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-161
Figure 7.8.44. San Vicente Dam: Spillway bridge joint separation of 20 cm at left support. MIDUVI
Inspection Commission photo (GPS: 2° 0' 2.3'' S, 80° 31' 54.9'' W).

Figure 7.8.45. San Vicente Dam: Spillway bridge joint separation of 7 cm at right support. MIDUVI
Inspection Commission photo (GPS: 2° 0' 1.36'' S, 80° 31' 54.7'' W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-162
7.9 Embankments
INTRODUCTION

During reconnaissance, the GEER-ATC team travelled on hundreds of kilometers of the


Ecuador roadway network, most of which lies on embankments. The majority of the highway
embankments performed as designed, maintaining serviceability of the road, and in some cases
requiring minor repair. This section presents observations of poor behavior of two bridge
approach embankments, whose failures impacted local connectivity, ability to quickly receive
supplies, and access of emergency vehicles. The two bridges are:

 Rio Chico (Las Chacras) Bridge, located approximately 8 km south of the town of
Rocafuerte (0° 58' 35.7'' S, 80° 25' 21.3' 'W), and

 Mejia Bridge, located about 8 km north of Portoviejo (0° 59' 22.7'' S, 80° 28' 11.1'' W).

EMBANKMENT FOUNDATION SUBSURFACE SOILS

The subsurface foundation soil conditions at the two bridge embankments provide the
bearing strata that control the stability and magnitude of deformations. Available geotechnical
information that can be used to determine the type and strength and compressibility of the
subsurface soil strata for studying the observed performance are provided in the next
paragraphs.

Rio Chico (Las Chacras) Bridge Embankment

A 2013 geotechnical report is available for the Rio Chico (Las Chacras) Bridge by LUP,
containing bridge foundation recommendations based on information from two borings at a
depth of 34 m below the surface, drilled at each abutment. Standard Penetration Testing (SPT)
was performed and selected soil samples were collected for laboratory tests. Both boring
results, shown on Fig. 7.9.1 and Table 7.9.1, indicate a subsurface of interbedded sand layers.

The SPT N-value in Boring 1 ranges from 2 to 9 bl/30cm in the top 7 m, increased to over
20 bl/30cm below 7 m to a depth of 32 m below ground. In Boring 2, SPT N-values range from
2 to 8 bl/30cm at the top 6 m and increase to 12 and 13 bl/30cm between a depth of 7 and 9 m.
The blow count SPT N-values increase to over 20 bl/30cm from a depth between 9 to the end
of the borings at 34 m. Table 7.9.1 was derived from Fig. 7.9.1, which describes the subsurface
profiles from the LUP Geotechnical Report of 2013.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-163
Figure 7.9.1. Soil classification from two borings at Rio Chico (Las Chacras) Bridge embankments
(after LUP Geotechnical Report, 2013).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-164
Table 7.9.1. Soil description from two borings at Rio Chico (Las Chacras) Bridge embankments.

Mejia Bridge Embankment

The subsurface soils at the Mejia Bridge, based on available borings and in-situ soil testing
from NYLIC (2001), are composed of layers of silts and clays as shown on Fig. 7.9.2. Boring
1 (right) shows a 3-m thick stiff crust underlain by a zone with undrained shear strength around
1 tons/m2 between 3 and 6 m, followed by strength at 2 tons/m2 up to a depth of 16 m, and then
generally increasing strength with depth.

Boring 2 (left) also shows a 3 m thick stiff crust layer followed by a zone with undrained
shear strength Su of 2 tons/m2 from 3 to 6 m depth, followed by Su of 4 tons/m2 to a depth of

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-165
20 m that is generally increasing with depth below that, with a stiffer zone approximately
between 18 and 24 m depth. The presence of the soft layer between 3-6 m could have
contributed to the instability, but also to potential long term time-dependent settlement of the
subsurface soils.

Zone of
Zone of
increased
increased
strength
strength

Figure 7.9.2. Undrained shear strength Su from two borings at the Mejia Bridge (NYLIC, 2001).

EMBANKMENT FILL

The weight of the fill used to construct the embankment generates the design load. The
strength and stiffness of the fill soil defines how steep the slopes can be constructed.

The embankments for the Rio Chico and Mejia Bridges appear to have been constructed
with soil that contains gravel, sands, and some fines, with unknown actual gradation (Fig.
7.9.3). The Mejia embankment incorporated use of geotextile and geogrids for the pavement
section design. These materials were there to improve the performance of the pavement and
not to reinforce the embankment. The material appears to be consistent throughout the fill.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-166
Geogrid

Geogrid
Geotextile

Figure 7.9.3. Embankment fill: Rio Chico (top left, GPS: 0°58'34.5"S, 80°25'20.0"W); Mejia (top right,
GPS: 0°59'25.4"S, 80°28'10.2"W) and its geotextile, geogrid (bot., GPS: 0°59'25.8"S, 80°28'10.5"W).

GEOMETRY AND FAILURE MODES

The embankment at the Rio Chico (Las Chacras) Bridge is approximately 12 m wide and
up to 5 m tall. The 12-m width includes two lanes and shoulders on each side. The embankment
at the Mejia bridge abutment is approximately 8 m in height and 25 m wide to accommodate 2
lanes of traffic in each direction, as well as 2 shoulders and a center gap. The slopes appear to
have been constructed at a 2H:1V slope (Horizontal to Vertical).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-167
In general, embankments can fail in various modes including circular planes, sliding
blocks, settlement, and lateral deformations as shown on Fig. 7.9.4 (Samtani & Nowatzki,
2006). The critical mode of failure depends on several factors, among which are the foundation
soil profile, depth, and strength; the static load magnitude and its distribution; and the geometry
of the embankment. In addition, seismic loads impose inertial forces and may soften and
decrease the strength of the bearing soils through cyclic loading or liquefaction.

Figure 7.9.4. Failure modes: circular (left) and sliding block (right) after Samtani & Nowatzki (2006).

The Rio Chico Bridge embankment experienced several failure modes. The reconnaissance
team observed large coherent cohesive blocks that settled, slid, rotated, and in several cases
moved in combination of these modes. Figure 7.9.5 shows the embankment settlement against
the bridge abutment punching through the bearing, measured between 30 and 60 cm relatively
to the bridge. Figures 7.9.6 and 7.9.7 were taken from opposite directions, revealing other
failure modes, such as a classic slope instability where one part of the embankment is rotating
and sliding away. The immediately adjacent block is rotating into the embankment in opposite
direction and the outside crest of the slope appears to plunge into the softened foundation soils.

Rio Chico Bridge

Settlement
30-60 cm
relative to bridge

Figure 7.9.5. Rio Chico embankment settlement (GPS: 0°58'34.78"S, 80°25'20.21"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-168
Sliding, rotating block
away from embankment
Crest of
slope plunging

Sliding, rotating block


into embankment

Figure 7.9.6. Rio Chico: Settlement and rotation of block in foreground and circular global failure in
background (GPS: 0° 58' 36.6" S, 80° 25' 23.7" W).

Crest of Sliding, rotating block


slope plunging into embankment

Sliding, rotating block


away from embankment

Figure 7.9.7. Rio Chico: Settlement and rotation of block in background and circular global failure in
foreground (GPS: 0°58'38.09"S, 80°25'25.67"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-169
Figure 7.9.8 shows a long section of road sliding and rotating into the Rio Chico
embankment as it plunges into the foundation soils. The long linear vertical crack along the
painted centerline of the embankment is likely the location of the pavement seam which could
have been leaking surface water into the embankment over time, causing weakness. Separation
of the pipe of about 20-30 cm, showing base sliding can be seen on Fig. 7.9.9. There is little
outward displacement expression along the outside slope face. Figure 7.9.10 shows depths of
cracks that reached 4 m, with consistent depth, regardless if the crack was 30 cm or 2 m wide.

Crest of
slope
plunging

Displacement
5 cm to 2 m

Sliding, rotation
away from
embankment

Figure 7.9.8. Rio Chico embankment: Block settlement and rotation (GPS: 0°58'41.2"S,
80°25'26.8"W). GEER member K. Rollins shown on background.

Separation of culvert
20 to 30 cm

Figure 7.9.9. Rio Chico embankment: Separation of pipe (GPS: 0°58'43.5"S, 80°25'26.7"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-170
3.7 m

2.1 m

Figure 7.9.10. Crack depth measurements along the Rio Chico embankment of 3.7 m (left, GPS:
0°58'72.0''S, 80°25'26.6''W) and 2.1 m (right, GPS: 0°58'34.9''S, 80°25'20.7''W).

At the Mejia Bridge the embankments appear to have failed in a global stability mode. This
was more evident at the south side where most of the damage to the bridge, walls, and
embankment were observed. The north side exhibited minor instability compared to the south.
The arrows of Fig. 7.9.11 show the direction of the observed ground movements.

Figure 7.9.11. Mejia Bridge and approaches plan view during construction. Arrows show the direction
of movement. Base image from Google Earth (©2016 DigitalGlobe).

An aerial view of the failure at the Mejia Bridge site (by COE-3) is presented on Fig. 7.9.12
with sketched markups of the observed movement of the embankment along assumed failure
planes. The failure planes appear to be either along the interface between the embankment and
the foundation soils (planar surface) or went through the foundation soils. The latter, indicative
of a more circular surface failure may be due to liquefaction-induced softening that allowed
the reduced-strength soil to shear in the earthquake. Damage at the SW approach embankment
with temporary access construction are shown on Figs. 7.9.13 and 7.9.14.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-171
Rotation of Sign
Lateral displacement
of toe of slope

Failure planes

Figure 7.9.12. Mejia Bridge: Aerial view of embankment failure with assumed movement and failure
modes marked with yellow (after COE-3, GPS: 0°59'24"S, 80°28'11"W).

Circular failure planes

Temporary repair
with new fill
to allow flow of traffic

Figure 7.9.13. Mejia Bridge embankment: View after earthquake (top, modified from Weiser-
Woodward, 2016); Repairs to approach to allow access (bot., GPS: 0°59'26.24"S, 80°28'10.19"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-172
Gabions displaced downward and
rotated away from embankment
similarly to slope movement
and perpendicular to roadway

Figure 7.9.14. Mejia Bridge: Failures of SW abutment and gabions (GPS: 0°59'23.5"S, 80°28'11.6"W).

The north abutment exhibited little damage to the walls and gabion slope adjacent to the
bridge, with some settlement and/or rotation of the slope next to the NE wing wall (Fig. 7.9.15).

Slope movement
along NE wing wall

No damage on
abutment, wing walls or gabions

Figure 7.9.15. Mejia Bridge: North abutment performance (GPS: 0°59'24.94"S, 80°28'11.26"W).

More details on the Mejia Bridge case study can be found in Section 7.7 of this chapter.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-173
7.10 Case Studies
This chapter presents observations and reconnaissance data and test results for cases of
essential facilities (hospitals, schools) and seismically isolated structures, and cemeteries.

IESS HOSPITAL, MANTA

The IESS hospital in southern Manta along the Avenida De La Cultura road was heavily
damaged during the main earthquake (Fig. 10.1.1), leaving the facility unable to function (see
Chapter 8 for structural observations).

Figure 7.10.1. IESS Hospital post-earthquake. GEER-ATC team visit on 4/28/16. Team member G.
Lyvers and X. Vera on the right (GPS: 0°57'18.5"S, 80°43'26.5"W).

During reconnaissance, on April 28th, the GEER team performed in-situ testing including
surface wave (MASW and MAM) and Horizontal-to-Vertical Spectral Ratio (HVSR). No
ground failures, ejecta or other evidence of liquefaction were observed at the site. The location
of the site and plan layout of the testing are shown on Figs. 7.10.2 and 7.10.3, respectively.
The MASW array is shown on Fig. 7.10.4 and the MAM array on Figs. 7.10. The HVSR setup
is shown on Fig. 7.10.6.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-174
SWM and HVSR
Testing Locations

Meters
0 1026

Figure 7.10.2. IESS Hospital location (GPS: 0°57'17.8"S, 80°43'25.8"W). Aerial photo pre-earthquake
(Digital Globe, ©2016).

Active MASW Array


Passive 2D MAM Array
Active Source Locations
Horz to Vert Spectral Ratio

Meters
0 85

Figure 7.10.3. IESS Hospital: Locations of active and passive surface wave arrays, and HVSR in-situ
testing performed by GEER on 4/28/16. Base image: Digital Globe (©2016).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-175
Figure 7.10.4. IESS Hospital: Active MASW array (GPS: 0°57'15.5"S, 80°43'25.8"W).

Figure 7.10.5. IESS Hospital: Passive MAM array (GPS: 0°57'16.0"S, 80°43'27.6"W).

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-176
Figure 7.10.6. IESS Hospital: HVSR testing (GPS: 0°57'15.9"S, 80°43'25.8"W).

Measured Vs values are presented in Table 7.10.1 with the dispersion curve and developed
Vs profile shown on Fig. 7.10.7 with. The site has a soft top layer with Vs of 178 m/s that
extends to a depth of 8.5 m below the ground surface, underlain by a layer with Vs of 288 m/s
extending to 51 m depth, where a weathered rock layer is encountered with Vs of 1,020 m/s.

Table 7.10.1 Median Vs values measured at IESS Hospital.

Layer Depth to Bottom Layer Shear Wave


Number of Layer Thickness Velocity
# (m) (m) (m/s)
1 8.4 8.4 176
2 51.0 42.5 287
3 Bedrock Bedrock 1020
Vs30 (m/s) 244 Vs to depth of 30 m
VsBR (m/s) 260 Vs to depth of bedrock
fsite (Hz) 1.28 Fundamental site frequency

This layering generated one major and one slight HVSR peak at the site, as shown on Fig.
7.10.8. The peaks at 1.44 Hz and 5.0 Hz match well with the theoretical peaks generated from
the median Vs profile developed at the site. Therefore, amplification effects are likely in these
two frequency ranges. Given that the IESS hospital is 5 stories tall, a rough estimate of the
natural frequency of the building 2.0 Hz (0.5 s, based on approximate structural period T of
0.1second per floor), which is close to the HVSR peak at 1.44 Hz. Therefore, soft soil
amplification likely increased the ground motion amplitude at the natural frequency of the
building, which may have contributed to the extensive damage observed at the site.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-177
700 0
Mean with +/- 1
Theoretical Fit
600
10

500
20
Phase Velocity (m/s)

400

Depth (m)
30

300

40
200

50
100

0 60
1 10 100 0 500 1000
Frequency (Hz) Shear wave Velocity (m/s)

Figure 7.10.7. IESS Hospital: Median theoretical Rayleigh wave dispersion curve for the top 1,000
profiles; fit to experimental dispersion data (left). Median of the top 1,000 Vs profiles obtained from
over 500,000 models searched during inversion (right).
4
1.44 Hz Experimental Mean HV
1.47 Hz
Experimental HV Peak
3 Theoretical HV
Theoretical HV Peak
H/V Spectral Ratio

0
0 1
10 10
Frequency (Hz)

Figure 7.10.8. IESS Hospital: Mean horizontal-to-vertical spectral ratio (HVSR). Red theoretical lines
represent the HVSR curve and peak for the median of the top 1,000 profiles.

RODRIGUEZ ZAMBRANO HOSPITAL, ISOLATED BUILDINGS, CEMETERIES

These sections will be included in the next version of this report.

GEER-ATC Muisne, Ecuador, Earthquake 7-Geotechnical Observations


Report Version 1 7-178
Blank Page
Blank Page
CHAPTER 8
Structural Observations

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
8.1 Building Inventory and Construction Types
This section presents the type of buildings in the area affected by the earthquake and
provides examples of observed buildings for each construction type.

I. Reinforced Concrete (RC) Buildings: These can be subdivided into three categories
based on the construction type.
a. Reinforced concrete buildings with 2-way joist waffle slabs, supported on
reinforced concrete columns. Typically form blocks in waffle slabs are not removed
after slab placement.

a) b)
Figure 8.1.1. Residential buildings under construction with 2-way waffle slabs a) Building in
Bahía de Caraquez. (GPS coordinates: 0°37’30.7”S, 80°25’31.5”W). b) Building in Manta
(GPS coordinates: 0°57’8.1”S, 80°42’26.9”W).

a)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-1
b)
Figure 8.1.2. a) Residential building under construction with 2-way waffle slabs in Tarqui,
Manta. (GPS coordinates: 0°57'28.67"S, 80°42'32.41"W). b) Details of reinforcement layout
in the back of the house.

b. Reinforced concrete buildings with 1-way beams and waffle slabs, supported on
reinforced concrete columns.

Figure 8.1.3. a) Residential building under construction with 1-way beams and waffle slabs in
Tarqui, Manta. (GPS coordinates: 0°57’27.3”S, 80°42’32.9”W).

c. Reinforced concrete buildings with moment resisting concrete frames and 2-way
waffle or flat slabs, supported on reinforced concrete columns.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-2
Figure 8.1.4. Rodriguez Zambrano Hospital building with moment resisting concrete frames
and waffle slabs in Manta. (GPS coordinates: 0°57’21.8”S, 80°44’30.8”W).

Figure 8.1.5. IESS Hospital building with moment resisting concrete frames and flat slabs in
Manta. (GPS coordinates: 0°57’18.0”S, 80°43’25.2”W).

Note: Residential reinforced concrete buildings are typically two to five stories in height
with an additional metal deck roof that covers the terrace. Several of these residences have 1
to 3 floors cantilevering more than 5 ft. The first floor of these residential buildings is generally
commercial. Other residential or commercial reinforced concrete buildings are typically 4 to 7
stories tall.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-3
II. Steel Framed Buildings: The reconnaissance team observed one steel framed
structure, which was the 911 ECU center.

Figure 8.1.6. The 911 ECU center in Portoviejo. It is a 3-story steel framed building (GPS
coordinates: 1°4’16.5”S, 80°26’47.5”W).

III. Wood Framed Buildings: There were wood framed 2-story buildings as well as 1-
story houses with roof metal decks.

Figure 8.1.7. Wood framed house with metal deck roof in Canoa (GPS coordinates:
0°27’47.7”S, 80°27’16.9”W).

Seismic Isolated Structures: There are some buildings, a bridge and a sculpture monument
that are base isolated in Ecuador. The isolation systems are friction pendulum or lead rubber

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-4
bearings. Some seismic isolated buildings include the Aerocity Sky Building in Guayaquil, a
new building under construction at ESPE in Quito and an Annex Building to the SOLCA
Hospital in Guayaquil. The Bahía to San Vicente Bridge “Los Caras” is seismic isolated with
a friction pendulum system; refer to 8.11.1 section for more information. The Guayas and Quil
sculpture steel framing structure installed on isolators in Guayaquil.

Figure 8.1.8. The lead rubber isolators under the Sky Building which is under construction in
Guayaquil. The reconnaissance team did not observe the structure. However, in the building’s
marketing website, it says it was visually inspected and no damages were observed after the
earthquake. (GPS coordinates: 2°8’39.7”S, 79°52’57.2”W - Photo obtained from
http://skybuilding.com.ec/ )

Figure 8.1.9. The Bahía to San Vicente Bridge is a seismic isolated bridge using triple friction
pendulum isolators. See section 8.10.1 for more information (GPS coordinates: 0°36’35.9”S,
80°25’25.6”W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-5
Figure 8.1.10. The Guayas and Quil monument (86 tons), which will be completed in July of
2016 in Guayaquil, is base isolated. (GPS coordinates: 2°9’16.7”S, 79°52’42.6”W - Photo
obtained from www.ecuavisa.com).

8.2 Construction Means/Methods/Implementation


INTRODUCTION

Past and current construction practices in Ecuador are one of the areas of concern related
to seismic vulnerability. The construction process, the quality of the materials, the lack of
seismic detailing and the inspection requirements during construction lead to the damage of
most of the buildings affected during the earthquake. The following sections are based on
observations and conversations with the residents of the areas visited.

EXISTING AND NEW CONSTRUCTION PRACTICES

Construction Workers

Several residences in the smaller towns of Ecuador were not built by construction
companies. Instead, individuals who have some knowledge of construction had built these
houses based on experience, without any form of design drawings. During the visits, there were
some laborers working on a damaged residence. When asked if they had construction drawings,
they said they did not have any available at the site.

Materials

Several existing reinforced concrete structures were built using aggregates containing salty
sand or salt water. This caused significant corrosion in the rebar which alters the strength of

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-6
rebar and the bond to the concrete. For example, in Manta which is adjacent to the Pacific
Ocean, observations in the field indicate that exposed rebar in structurally damaged buildings
was previously corroded in some cases, with an effective reduction in diameter of the
longitudinal, as well as the transverse steel.

(a)

(b)
Figure 8.2.1. Building materials near Los Esteros Market in Tarqui, Manta. a) Fine aggregate
with sea shells. b) Good quality coarse aggregate (GPS coordinates: 0°57'8.94"S,
80°42'29.38"W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-7
a)

b)
Figure 8.2.2. Signs of severe corrosion on reinforcing steel previous to the earthquake. a)
Collapsed electric pole with signs of reinforcement corrosion and reduction of effective area
of longitudinal steel in Tarqui, Manta (GPS coordinates: 0°57'29.92"S, 80°42'36.45"W). b)
Signs of severe corrosion in the steel at the base of a 1 st story column of a 3-story building
tagged for demolition in Tarqui, Manta (GPS coordinates: 0°57'9.04"S, 80°42'41.54"W).

In some existing buildings, there was some rebar sticking out from the top of the columns,
probably so that a floor could be added to the building in the future. However, leaving the rebar
exposed for long periods of time also causes the rebar to corrode.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-8
In waffle slab construction, there are several buildings that have the original form blocks
as part of the slab.

Figure 8.2.3. New construction where the form blocks have been left in the waffle slab in
Tarqui, Manta (GPS coordinates: 0°57’27.3”S, 80°42’32.9”W).

Reinforced Concrete Detailing

There is a lack of seismic detailing as well as typical gravity detailing in existing and new
construction. This was observed in the majority of the buildings. Dowels were missing in slab
and column connections. The hoops were made with smooth steel with approximate diameter
of 1/4 in. [6.4 mm] in most cases. The stirrup spacing seemed to be as large as the effective
element depth for beams and columns. Also, the stirrup spacing in 2-way slabs was large.
Additionally, columns were not confined; stirrups spacing did not decrease towards the ends
of the columns. Smooth longitudinal steel was observed in columns of some older multistory
moment frame buildings.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-9
a)

b)
Figure 8.2.4. a) A 1st floor column in the 3-story Hotel Las Gaviotas in Tarqui, Manta. There
is significant spalling associated to both, previously corroded steel and the seismic demand
imposed by the earthquake (GPS coordinates: 0°57’5.7”S, 80°42’48.5”W). b) Small-diameter-
hoop spacing details of a 1 st floor unconfined column in a 4-story commercial building in
Tarqui, Manta (GPS coordinates: 0°57’14.9”S, 80°42’53.9”W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-10
Figure 8.2.5. Damage at the base of 6-story mixed-occupancy building in Tarqui, Manta. There
is significant spalling at the bottom of the first floor columns, showing smooth longitudinal
rebar and hoop spacing of approximately 2/3 of the column width. (GPS coordinates:
0°57'14.55"S, 80°42'53”W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-11
a)

b)
Figure 8.2.6. a) Failure of first floor columns in building under construction in Canoa. The
column has large stirrup spacing. b) The slab to column connection at the edge shows a lack of
dowels, therefore providing no structural integrity (GPS coordinates: 0°27'42.67"S,
80°27'20.67"W)

a)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-12
b)
Figure 8.2.7. Slab reinforcement detail in a residential building under construction in Tarqui,
Manta. a) General view of joist, beams, and slab reinforcement; b) Details of beam
reinforcement near a beam-column joint (GPS coordinates: 0°57'28.67"S, 80°42'32.41"W).

Evidences of good transverse reinforcement detailing were scarce. Columns in multistory


RC building are subjected to combined flexural-tension and flexural-compression actions due
to the demand imposed by earthquake loading. Proper transverse/seismic longitudinal and
transverse reinforcement layouts in columns must ensure the protection of the concrete core
when the concrete cover spalls-off due to the aforementioned actions. Such proper
reinforcement layout generally comprises close-spaced hoops and crossties, both with seismic
135-degree hooks anchored into the core, along with a dense repartition of longitudinal steel
in the perimeter of the column. This creates a tight reinforcing cage that guarantees the concrete
core confinement, therefore ensuring adequate ductility of the structural element. Figure 8.2.8
shows a column in Zona Cero of Portoviejo, showing an apparently healthy core thanks to
proper longitudinal and transverse reinforcement detailing.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-13
Figure 8.2.8. Column with spalled-off concrete cover and evidence of healthy a core thanks to
closely spaced perimeter hoops and cross-ties. Zero Zone in Portoviejo. (GPS coordinates:
1°3'13.53"S, 80°27'8.60"W)

Inspections

In current practices, for both public and private construction, there is an engineer of record.
In the present law in Ecuador, for public building construction, it is obligatory to continuously
audit the construction. An engineer specially contracted or an engineer part of the staff of the
public office to whom construction belongs holds inspections, and accepts or rejects practices
as per the contract and construction drawings as part of an audit. However, for private
constructions, there are no mandatory inspections and it becomes the sole responsibility of the
engineer of record, but in large private constructions the owner contracts an engineer to audit
the project.

8.3 Building Codes and Evolution


Following major earthquakes in the 20 th century including the 1906 earthquake in
Esmeraldas, the 1979 earthquake in Tumaco and the 1987 earthquake in Napo, the Ecuador
building codes have developed over time to provide improved seismic resistant design
guidelines and regulations. This section provides a summary of the seismic design chapter of
the CEC-1977, 2001 and of the current NEC-2015 building code of Ecuador.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-14
CEC-1977 Ecuadorian Building Code

In this code, the minimum base shear was computed as

V = ZIKCSW (8.3.1)

where Z was a seismic factor which was equal to 1.0, 0.75 or 0.5 depending on the seismic
zone in which the structure was located, I was an importance factor which varied between 1
and 1.5, K was a factor which allowed reduction of lateral forces as a function of the structural
system and was equal to 0.67 for reinforced concrete frames, W is the total dead load of the
structure and C and S are factors which took into account the variation of forces as a function
of the fundamental period of vibration and of the local site conditions and were given by

1
C= (8.3.2)
15 T

æT ö
2
T T
S =1.0 + - 0.5çç ÷÷ when £1 (8.3.3)
Tg è Tg ø Tg

æT ö
2
T T
S =1.2 + 0.6 - 0.3çç ÷÷ when >1 (8.3.4)
Tg è Tg ø Tg

where T is the fundamental period of vibration of the structure and Tg is the characteristic

period of the site. The code specified that when Tg was unknown then the soil factor S needed

to be taken as equal to 1.5.

Equation 8.3.1 in this version of the code was the same as that used in the 1976 Uniform

Building Code.

CEP-INEN-2001 Ecuadorian Building Code

The 2001 version of the code included an improved seismic zonation of the country based
on a new probabilistic seismic hazard analysis. In addition to equivalent static analysis it
introduced modal response spectrum analysis. In this code, the equation to compute the

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-15
minimum base shear is similar to the 1994 UBC in the United States and was computed with
the following equation

ZIC
V= W (8.3.5)
RF pFe

where Z was a seismic factor which varied between 0.15 and 0.4 depending on the seismic
zone in which the structure was located as indicated in figure 8.3.1 and table 8.3.1. With
exception of a few counties in the east, provinces of Esmeraldas and Manabi were located in
zone IV with a seismic factor Z=0.4, which is the same value that was used in the regions of
highest seismic hazard in the US such as California. The importance factor was equal to 1.5
for hospitals and clinics, military installations, fire and police stations, air control towers, etc.
A factor I=1.3 was specified for schools, museums, stadiums and churches that could hold
more than 300 people as well as any structure that could hold more than 5,000 people of any
public building that provided continuous service. It is interesting to point out that in the UBC
1994 hospitals only required an importance factor of 1.25.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-16
Figure 8.3.1 Seismic zonation of Ecuador in the CPE INEN CEC-2001.

Table 8.3.1 Factors Z as a function of the seismic zones in the country

Seismic Zone I II III IV


Factor Z 0.15 0.25 0.30 0.40

The factor C, similarly to the one in the 1977 version of the code, was a factor to consider
the variation of seismic coefficients as a function of the period of vibration of the structure,
were given by

1.25S S
C= (8.3.6)
T

R is a strength reduction factor that was 12 for moment resisting frames either in reinforced
concrete or structural steel with reinforced concrete shear walls (dual systems) and equal to 10
for moment frames or columns and flat slab systems with RC shear walls. In the code flat slabs
are referred to as frames with embedded beams. For flat slabs and columns the code allowed
an R=9. For confined masonry structures an R of 5 was specified. These reduction factors are
based on forces calibrated to allowable stress so they are similar to RW factors in 1994 UBC.
Factors P and E were factors that varied between 0.8 and 1.0 to reduce the reduction factor

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-17
when there are irregularities in plan (P) or in elevation (E) in the structure. Factors S depend
on site conditions as indicated in table 8.3.2.

Table 8.3.2 Factors S as a function of site conditions.

Soil
Description S
profile
S1 Rock or firm soil 1.0
S2 Intermediate soils 1.2
S3 Deep soil soils 1.5
S4 Special soil conditions 2.0

Lateral deformations are computed as M=RE where M are the estimated inelastic
deformation and E are the lateral deformation demands computed with a linear elastic analysis
with reduced forces. The maximum allowable drift for reinforced concrete structures was 0.02
and 0.01 for masonry structures.

Figure 8.3.2 Design spectra in the CEC-2001.

NEC-2015 Ecuadorian Building Code

The Norma Ecuatoriana de la Construcción (NEC). NEC is promoted by the Habitat and
Human Settlements of the Ministry of Urban Development and Housing (MIDUVI). The NEC-
2015 building code replaces the NEC-2001 This is a modern code based on recent codes and
design guidelines such as ASCE 7-10, ASCE 41, FEMA 440, etc. It has the following ten
chapters.
1. Loads (non-seismic)
2. Seismic Loads and Seismic Resistant Design
3. Seismic Rehabilitation of Structures

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-18
4. Reinforced Concrete Structures
5. Structural-Masonry Structures
6. Geotechnical and Foundations
7. Steel Structures
8. Wood Structures
9. Glass
10. Houses of Up to Two Floors With Spans Up to Five Meters

The 2015 version of the code divides the country into 5 zones with seismic coefficients
varying from 0.15 to 0.6 which corresponds to short period spectral ordinates with a probability
of exceedance of 10% in 50 years (return period of 475 years). Design spectra an equations
defining the three spectral regions are shown in Figure 8.3.4.

Figure 8.3.3 Seismic zonation for Ecuador in the 2015 Ecuador Building Code.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-19
Figure 8.3.4 Design spectrum in the NEC-2015.

The factor  which represents the amplification of acceleration demands from PGA (i.e.,
T=0s) to T=0.1s is equal to 1.8 in the coastal regions of the country. Exponent r which defines
the descending branch is equal to 1.0 for structures in site classes A, B and C and equal to 1.5
for site classes D and E. Factors Fa and Fd are soils modification factors in the short
(acceleration) and displacement spectral regions, respectively. These factors are somewhat
analogous to factors Fa and Fv is recent seismic codes in the U.S., except that in the Ecuador
building code there is a third modification factor due to soil conditions Fs which account for
nonlinearity in the soil in the case of strong ground motion intensities.

The following figure shows the spectral displacement curve. The curve only applies for
structures with periods less than or equal to T L.

Figure 8.3.5 Displacement spectrum in the CEC-2015.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-20
The code requires the use of a vertical component equal to two thirds of the intensity of the
horizontal component. Importance factors remained unchanged to those in the 2001 code
requiring hospitals and other essential facilities to be designed with an importance factors of
1.5.
The code permits three design methods: (1) Force-based procedure; (2) Displacement-
Based procedure using equivalent linearization and (3) plastic design combined with capacity
design. In the force based procedure the R reductions factors are equal to 8 for dual systems
and equal to 8 for moment frames.
Lateral deformations are computed as M=0.75RE where M are the estimated inelastic
deformation and E are the lateral deformation demands computed with a linear elastic analysis
with reduced forces. The maximum allowable drift are the same as those in the 2011 version
of the code in which the maximum drift ratio for reinforced concrete structures is 0.02 and 0.01
for masonry structures.

8.4 Damage Assessment/Tagging/Incorporation of ATC-20

The ATC-20 report, Procedures for Post-earthquake Safety Evaluation of Buildings,


presents procedures and guidelines for the safety evaluation of buildings. This report was
published by the Applied Technology Council (ATC) in 1989 and updated in 2005. The
evaluations are made to determine whether buildings are safe for continued use, or if entry
should be restricted or prohibited. The outcome of the evaluation is communicated via a
placard that is posted on the building. Each building can be tagged into three different
categories, as described below (ATC, 2005). - Green: The structure has been inspected and no
apparent structural hazard has been found; no restrictions are placed on use or occupancy. -
Yellow: This structure has been inspected and a hazardous condition exists (or is believed to
exist) that requires restrictions on the occupancy or use of the structure; entry and use are
restricted as indicated on the placard. Red: This structure has been inspected, found to be
seriously damaged and is unsafe to occupy; entrance into the building is restricted.
Following the earthquake, the Ministry of Urban and Housing Development (MIDUVI) in
Ecuador adopted a Spanish translation of selected ATC-20 documents and placards for use by
inspectors. This translation was made by the Universidad de San Francisco de Quito (USFQ)
as part of a MS thesis aimed at exploring the potential of using intellectual property within
ATC-20 for assessing post-earthquake conditions in Latin America. In addition, MIDUVI also

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-21
utilized a smart phone application developed for Android and Apple IOS mobile devices based
on the translated and adapted forms, allowing the inspection results to be uploaded directly to
the database on the principal server for daily reports.
The adapted Spanish translation of ATC-20 also took into account the types of buildings
that were damaged during this earthquake. The original ATC-20 methodology restricts entry
and occupancy to buildings with a yellow tag (as indicated on the placard). However, for the
Ecuador building evaluations, MIDUVI enlisted a change such that yellow tagged buildings
were occupiable with some restrictions.
In some cities, zones, called Zone Zero, were established, in which only military personnel
or individuals with government permits were allowed to enter. These areas were closed off to
public for security reasons. The reconnaissance team had access to these areas.

Figure 8.4.1. Green tag used to evaluate post-earthquake safety. It says that the structure has
been inspected and no structural damages or risks are apparent. Tags were provided by
MIDUVI.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-22
Figure 8.4.2. Yellow tag used to evaluate post-earthquake safety. It says that the structure has
been inspected. The structural engineer would list all of the damages as well as the areas that
are restricted. Tags were provided by MIDUVI.

Figure 8.4.3. Red tag used to evaluate post-earthquake safety. It says that the structure has
been inspected and there is structural damage and risk. Entering the building is restricted unless
a person brings a legal document saying otherwise. It warns that entering the property could
cause injuries or even death.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-23
Figure 8.4.4. Damage assessment form used by structural engineering professionals to record
the structural damage of the property.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-24
The following damage survey shows the percentage of structures in each tag category
evaluated by April 29, 2016. The areas with the most damage are the cities of Manta,
Portoviejo, Muisne and Bahía de Caráquez.

Figure 8.4.5. This shows the damage assessment in the affected cities until April 29, 2016.
The survey was completed in the zone 0 and in the most affected areas. The damage survey
and chart were provided by COE3, MIDUVI.

The following maps show the color tag designation of different properties after being
evaluated for damage and risk.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-25
Figure 8.4.6. Map of colored tags of properties in Manta after post-earthquake safety
evaluation. The map was provided by COE3, MIDUVI.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-26
Figure 8.4.7. Map of colored tags of properties in Manta after post-earthquake safety
evaluation. Information updated until April 29, 2016. The map was provided by COE3,
MIDUVI.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-27
The reconnaissance team compared the damage assessment information until April 29 th in
the ECU M7.8 earthquake provided by the MIDUVI to results with the U.S. Code and the Chile
Maule Earthquake. The data was obtained from ATC 63 and Chris Poland ASCE 31/41
webinar.

Figure 8.4.8. Probability of Failure. Comparison between US Code, Chile Maule Earthquake
and ECU M7.8 Ecuador Earthquake updated data until April 29th. Data from ATC 63, Chris
Poland ASCE31/41 webinar and COE3, MIDUVI Ecuador.

8.5 Main Structural Observations and Structural Behavior

INTRODUCTION

During the seismic event that occurred on April 16 th, 2016 (Mw 7.8), 29 kilometers south-
southeast of Muisne, numerous buildings and houses in the northwest of Ecuador were
damaged. Soon after, several aftershocks followed. Major aftershocks include: the two
earthquakes on April 20th (Mw 6.2 and 6.0), 19 km west-northwest and 12 km north-northeast
of Muisne, respectively; the earthquake on April 22 nd (Mw 6.0), 34 km north-northwest of
Bahía de Caráquez; and the two earthquakes on May 18 th (Mw 6.7 and 6.8), 34 km west-
northwest and 24 km northwest of Rosa Zárate, respectively (three weeks after the
reconnaissance observations).
Most of the structural damage was caused during the 1 st major earthquake. In general, all levels
of damage were observed, from small cracks to complete collapse of buildings. The damage at
each building was dependent on the structural system, poor seismic detailing, poor construction

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-28
practices, quality of materials, and intensity of the ground motion. The most damage occurred
in buildings in Pedernales (~35 km from the epicenter), Canoa (~58 km from the epicenter),
Portoviejo (~169 km from the epicenter) and Manta (~172 km from the epicenter) based on the
places visited by the reconnaissance team.
MAIN OBSERVATIONS

Generally there was medium to severe damage. In residential areas along the coast,
buildings were significantly damaged. The following categories characterize the typical level
of damage in the commercial and residential buildings.

Failure of non-structural walls:

Most interior and exterior walls were composed of masonry, bricks, hollow clay or concrete
blocks. These walls were most likely conceived and constructed as non-structural elements.
However, during the earthquake these walls were subjected to seismic loads and functioned as
part of the lateral load resisting system thanks to their large in-plane stiffness. The walls were
not detailed to resist such loads and failed. Most walls were cracked or in severe cases collapsed
out-of-plane. There were no concrete shear walls observed. The collapse of these non-structural
walls was the most life threating failure, apart from complete structural collapse.
Some building codes in other countries in the region, such as the Reglamento Colombiano
de Construcción Sismo Resistente NSR-10, mandatory in Colombia, recommend either tying
the non-structural infills to the surrounding structure (typically beam-column framing system)
if compatibility of deformations can be guaranteed, or fully separating them laterally such that
the surrounding structure can freely deform without interaction with the non-structural
elements. In this latter case, the non-structural component has to be anchored from the bottom
or hung from the top and must be able to sustain the inertial forces without failure.
In the current Ecuadorian building code, Norma Ecuatoriana de la Construcción (NEC)
(NEC, 2014), there is a chapter focused on structural masonry (Mamposteria Estructural in
Spanish). It clearly states that the structural masonry should be identified and specified in
structural drawings. It lists all of the masonry specifications and details that should be included
in the drawings.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-29
Figure 8.5.1. Residential building in Tarqui, Manta, where the exterior walls collapsed. The
roof also fell in this property. (GPS coordinates: 0°57’14.6”S, 80°42’38.5”W).

a)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-30
b) c)

Figure 8.5.2. a) Collapsed brick walls in building in IESS Hospital in Manta (GPS coordinates:
0° 57' 18.0"S, 80° 43' 26.47"W). b) Concrete building with wood and metal deck extension in
Portoviejo. Non-structural wall completely collapsed (GPS coordinates: 0°57’14.6”S,
80°42’38.5”W). c) Concrete building in Canoa where brick and masonry exterior walls
collapsed. (GPS coordinates: 0°27'46.59"S, 80°27'23.56"W).

Figure 8.5.3. This is a site of new construction. These are dowels at the concrete columns to
be attached to the non-structural walls (GPS coordinates: 0°57'7.31"S, 80°42'26.44"W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-31
Soft story damage

Various buildings had a commercial 1 st floor with residences on the upper floors. Several
of these had soft story damages where one or two floors had been completely crushed. Also,
there was significant hinging of the 1 st floor columns in some of these damaged structures. This
could have been contributed by the change in stiffness at the compromised story, which was
typically surrounded (from above, or above and below) by stiffer stories with larger masonry
infills density.

a)

b) c)
Figure 8.5.4. Soft story failure: a) The first floor of this building in Manta collapsed. Some of
the rebar in the columns that failed was corroded. Additionally, the columns were not confined
(GPS coordinates: 0°57’14.8”S, 80°42’31.6”W). b) The first floor of this building in Tarqui,
Manta collapsed. Column confinement was not apparent (GPS coordinates: 0°57’29.9”S,
80°42’42.1”W). c) The second floor of this building in Portoviejo collapsed (GPS coordinates:
0°57’29.9”S, 80°42’42.1”W).
GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations
Report Version 8-32
Figure 8.5.5. Plastic hinging in the 1st floor columns of this residential building with 1 st floor
commercial occupancy in Tarqui, Manta (GPS coordinates: 0°57’7.1”S, 80°42’55.3”W).

Figure 8.5.6. Plastic hinges formed in the 1 st floor columns of this building in Manta (GPS
coordinates: 0°57’9.5”S, 80°42’49.4”W).

Captive columns

For some buildings, the stiff partial-height non-structural masonry walls or concrete blocks
between columns restrained the column height. This exacerbates the shear demand on the
column when plastic flexural demand concentrates at their ends.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-33
a) b)
Figure 8.5.7. a) The 4th and 5th floors collapsed in this residential building in Portoviejo (GPS
coordinates: 1°3'13.53"S, 80°27'8.60"W). b) The concrete blocks restrained the column height,
causing captive column failure for this column.

Rebar corrosion

There was significant corrosion observed in the rebar in reinforced concrete structures. This
was caused by the use of salty sand and/or salty water in the concrete admixture. Also, it could
have been caused by the exposure of the rebar for a long time prior to pouring concrete.
Observations in the field indicate that, for some structures, available bars of longitudinal and
transverse reinforcement were reduced in cross section area due to corrosion.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-34
a)

b)
Figure 8.5.8. a) Rebar corrosion in a beam-column connection, and slab reinforcement in a 3-
story building located in Tarqui, Manta (GPS coordinates: 0°57'7.94"S, 80°42'41.90"W). b)
Rebar area reduction at the base of a column in a 2-story house located in Tarqui, Manta (GPS
coordinates: 0°57'19.77"S, 80°42'40.68"W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-35
Pounding

In highly populated areas such as Zona Cero (Zero Zone) in Manta and Portoviejo,
separation between adjacent structures was not usual. This facilitated unfavorable interaction
between structures that had no evident seismic gap. During the earthquake, the building drifts
caused them to pound each other which caused severe damage in columns, and/or caused
permanent displacement for some buildings.

a) b)

c)
Figure 8.5.9. a) Pounding of independent buildings at the Bahia Airport (GPS coordinates:
0°36’5.3”S, 80°24’27.0”W). b) Pounding of two buildings in Portoviejo. The reconnaissance
team observed that the roof metal deck is continuous at the separation (GPS coordinates:
1°3’11.6”S, 80°27’8 c) Pounding evidence in Hotel Las Gaviotas in Tarqui, Manta (GPS
coordinates: 0°57'6.08"S, 80°42'48.85"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-36
Structural damage due to liquefaction and soil failure

Substantial settlement was observed in some houses in Tarqui, Manta due to liquefaction
and soil failure. This settlement caused some buildings to sink significantly.

a) b)
Figure 8.5.10. A house that settled more than a foot in Manta. a) Evidence of the settlement
with respect to an adjacent house. b) Interior of the house (GPS coordinates: 0°57'15.5"S,
80°42'33.7"W)

Figure 8.5.11. House settlements and soil failure evidence in Tarqui, Manta. (GPS coordinates:
0°57'18.0"S, 80°42'38.9"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-37
BUILDING COLLAPSE
There were several buildings that completely collapsed. This section shows a few of these
catastrophic failures.

Figure 8.5.12. Sideway collapse of residential structure in Pedernales (GPS coordinates:


0°4'4.8"N, 80°3'23.5"W).

Figure 8.5.13. Building collapse in Pedernales. The columns shown have hinges at both ends
(GPS coordinates: 0°4'3.94"N, 80°3'24.6"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-38
Figure 8.5.14. Collapse of building with flat slab construction in Pedernales (GPS coordinates:
0°4'2.5"N, 80°3'33.9"W)

Figure 8.5.15. Collapse of the Unidad Educativa “Linus Pauling” building in Manta (GPS
coordinates: 0°57'24.4"S, 80°42'40.8"W). See section 8.8 for more information regarding this
structure.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-39
Figure 8.5.16. Building that was under construction collapsed. The construction type was
reinforced concrete 2-way flat slabs supported on columns (GPS coordinates: 0°27'43.8"S,
80°27'21.0"W).

8.6 Non-structural Components

INTRODUCTION
Damage of non-structural components was extensive and significant in the meizoseismal
area. This section presents observations of damage to non-structural components in buildings
that did and did not have damage and buildings for which the non-structural components were
an important factor in the collapse.

These observations were made during reconnaissance efforts in the period from April 26 th
to May 2 nd, 2016. The impact of this damage was substantial in the function and the economy
of the cities. Non-structural damage caused businesses to stop operating, including banks,
restaurants, and stores, and the closing of hospitals and an airport terminal.

NON-STRUCTURAL CATEGORIES AND PERFORMANCE


We have grouped the non-structural components observed during reconnaissance in three
categories, as per the commonly used references of FEMA E-74 and ASCE7-05/10:
 Architectural
 Building utility systems (Mechanical, Electrical, and Plumbing)
 Furniture and contents
During reconnaissance, both acceleration and displacement or driftsensitive non-
structural components were inspected. Accelerationsensitive components include suspended

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-40
ceilings, lights, mechanical equipment, etc. Drift-sensitive components include partitions,
façades, etc. Bracing and anchorage to structural elements is important in the behavior of the
non-structural elements. The presence, type, and general conditions of the bracing or anchorage
were documented by the reconnaissance teams. Specific observations with possible
explanations for selected non-structural components in each category are presented in the
following paragraphs

ARCHITECTURAL COMPONENTS
The following architectural components and potential causes of damage are discussed and
illustrated in Figures 8.6.1 through 8.6.12: egress means (stairs); freestanding frames; exterior
walls; façade; glazing; interior partitions; lighting; suspended ceilings; unreinforced masonry;
and veneers.

Egress Means – Stairs:


As the primary means of egress, the performance of stairwells is critical following an
earthquake. Observations of damage in stairways often involved lack of rolling supports at
either end to accommodate interstory drift.

Figure 8.6.1. Concrete stair damaged and blocked by fallen infill walls at the IESS Hospital,
Manta. (GPS coordinates: 0°57'17.8''S, 80°43'25.4''W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-41
Figure 8.6.2. Concrete stairs damaged in Residential houses, Pedernales. (GPS coordinates:
0°4'12.9''N, 80°3'18.5''W).

Figure 8.6.3. Concrete stairs damaged, building in construction, Portoviejo. (GPS coordinates:
0°59'48.42'S, 80°28'0.4''W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-42
Glazing:
Façade and overhead glazing, although typically made with safety glass, were not designed

for seismic drift. FEMA (E-74) guidelines for non-structural components provides

recommendations for adding more space around the pane of glass where it is mounted between

stops or molding strips in order to accommodate greater frame distortions without cracking the

glass.

Figure 8.6.4. Glazing damaged at the IESS Hospital, Manta. (GPS coordinates: 0°57'17.8''S,
80°43'25.4''W)

Figure 8.6.5. Glazing damaged at residential houses, Pedernales (GPS 0°4'14.5''N,


80°3'15.3''W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-43
Exterior Walls and Interior Partitions:
Walls in Ecuador are primarily made with concrete block, particularly in buildings taller
than 3 stories. Also common but observed with less frequency are hollow clay tile. In both
cases there are unreinforced and typically are not attached to the structure to prevent out of
plane failure. Figure 8.6.6 to 8.6.8 show different examples of heavy partitions that failed and
fell during the event. Figure 8.6.9 shows a typical method for mitigating damage to partitions
by installing angles connected to the slab (FEMA). Falling masonry in hallways and stairwells
is a particular hazard for occupants attempting to exit buildings during an earthquake.

Figure 8.6.6. Examples of commonly observed failure in exterior masonry infill walls.
Universidad Técnica de Manabi in Portoviejo (GPS coordinates: 1°2'39''S, 80°27'21''W).

Figure 8.6.7. out of plane failure of exterior masonry infills at IESS Hospital, Manta (GPS
coordinates: 0°57'17.8''S, 80°43'25.4''W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-44
Figure 8.6.8. In plane failure of interior hollow clay masonry partitions that then led to
out-of-plane failures at IESS Hospital, Manta (GPS coordinates: 0°57'17.8''S, 80°43'25.4''W)

Figure 8.6.9. Cantilever wall damaged at “Fuerte Militar Manabi” Portoviejo (GPS
coordinates: 1°3'6.6''S, 80°28'30.8''W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-45
Figure 8.6.10. Partition damaged at Eloy Alfaro International Airport, Manta (GPS
coordinates: 0°57'11.34''S, 80°41'3.4''W). Method for mitigating damage to partitions by
installing angles connected to the slab (FEMA E-74).

Lighting:
Recessed light fixtures did not have positive support from hanging, or any special safety

devices for the attachment of lens covers. Figure 8.6.8 shows damage to recessed lighting in

the Eloy Alfaro International Airport, Manta.

Street lights poles along public sidewalks failed at the base, the supporting precast columns

lacked sufficient strength to withstand the earthquake movements. In some cases the

reinforcement was corroded and the post broke at the base. Figure 8.6.7 shows damage to lamp

posts in the Tarqui, Manta area.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-46
Figure 8.6.11. Precast concrete lamp posts damaged at Tarqui, Manta (GPS coordinates:
0°57'27.7''S, 80°42'35''W)

Suspended Ceilings:
There was no bracing of the suspension grid for acoustic lay-in tile or gypsum board
ceilings. Figure 8.6.12 shows pieces of ceiling that fell from the ceiling the Eloy Alfaro
International Airport, Manta (0°57'11.34''S, 80°41'3.4''W), similar to damage recorded at
several locations in the meizoseismal Manabí area. Figure 8.6.14 illustrates a mitigation option
recommended by FEMA to install diagonal bracing for ceiling supports.

Figure 8.6.12. Recessed light fixtures and ceiling damaged at Eloy Alfaro International
Airport, Manta (GPS coordinates: 0°57'11.34''S, 80°41'3.4''W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-47
Figure 8.6.13. In plane failure of interior hollow clay masonry partitions that then led to out-
of-plane failures at IESS Hospital, Manta (GPS coordinates: 0°57'17.8''S, 80°43'25.4''W)

Figure 8.6.14. Mitigation option to provide diagonal bracing for ceiling systems in FEMA E-
74.

Roof parapets:

Significant damage to parapets due to lack of bracing and pounding with adjacent buildings

during the earthquake events was observed in the area as shown on Figure 8.6.15 to 8.6.16

They represent a particular hazard for pedestrians and occupants attempting to exit damaged

buildings.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-48
Figure 8.6.15. Parapet damaged at the Bahia de Caraquez “Los Perales” Airport, Bahia de
Caraquez. And truck damaged by the parapet piece. (GPS coordinates: 0°36'5.2''S,
80°24'29''W).

Figure 8.6.16. Parapet damaged at the Jaramijo Port Building, Jaramijo. (GPS coordinates:
0°56'38.6''S, 80°38'16.1''W) and Unidad educative San Jose (GPS coordinates: 0°57’14.9''S,
80°42'45''W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-49
Veneers:

Adhered or anchored veneers experienced damage due to the deformation of the backing
substrate, heavy partitions. Figure 8.6.17 shows the tile veneer at the elevator shaft of the IESS
Hospital, Manta.

Figure 8.6.17. Veneer damaged at the IESS Hospital, Manta. (GPS coordinates: 0°57'17.8''S,
80°43'25.4''W).

BUILDING UTILITY SYSTEMS

Observations were made in MEP (Mechanical Electrical Plumbing) building utility


systems, namely piping; lifeline infrastructure; equipment; and electrical panels. Detailed
observations on lifeline infrastructure of potable and wastewater networks are presented in
Chapter 9. Characteristic damage in the remaining systems is illustrated in Figures 8.6.16
through 8.6.17 and discussed below.

Piping and ducts:


There was no bracing of the suspension of the electrical pipes, HVAC ducts and waste pipes.
Figure 8.7.18 shows pipes and ducts that fell from the ceiling at IESS Hospital, Manta
(0°57'17.9''S, 80°43'26.37''W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-50
Figure 8.6.18. Pipes and ducts fallen from the ceiling at IESS Hospital, Manta. (GPS
coordinates: 0°57'17.9''S, 80°43'26.37''W).

Equipment:
Heating, ventilating, and air conditioning (HVAC) equipment was typically not properly
braced or anchored. Large air conditioning (AC) units are typically placed on building roofs.
When poorly supported by simply sitting on steel stands, the AC units displaced or fell over,
as shown on Figure 8.6.19. Figure 8.6.20 shows appropriate bracing and bolting to the exterior
wall of HVAC units in Tarqui, Manta that suffered no damage.

Figure 8.6.19. HVAC equipment on the roof IESS Hospital, Manta. (GPS coordinates:
0°57'17.9''S, 80°43'26.37''W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-51
Figure 8.6.21. AC unit attachment at the residential area, Manta. (GPS coordinates:
0°57'29.3''S, 80°42'36.4''W).

FURNITURE AND CONTENTS

Desktop computers, printers and other accessories were not anchored or tethered. Figure
8.6.22 shows how many of these items were displaced or fell onto one another, photo provided
by the HRZ Hospital Communication Manager to GEER.

Figure 8.6.22. Furniture displacement at the HRZ Hospital, Manta. April 17 th. (GPS
coordinates: 0°57'15.3''S, 80°44'32.8''W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-52
Heavy furniture experienced rotational movement due to lack of floor connections. File
cabinets at the HRZ Hospital fell due to the lack of anchorage. When the GEER reconnaissance
team visited the site, the hospital workers were organizing the files and anchoring the shelves
to the wall and between them. Figure 8.6.23 and 8.6.24 show the shelves, the new anchors and
the photo of the file room on April 17, one day after the earthquake, provided by the HRZ
Hospital Communication Manager to GEER.

The special light fixtures in the surgery rooms show no damage in the anchorage. Oxygen
tubes were chained to the wall as shown in Figure 8.6.25.

Figure 8.6.23. File cabinets at the HRZ Hospital, Manta. April 17 th. (GPS coordinates:
0°57'15.3''S, 80°44'32.8''W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-53
Figure 8.6.24. File cabinets at the HRZ Hospital, Manta. April 28 th. Bracing to walls and
between cabinets. (GPS coordinates: 0°57'15.3''S, 80°44'32.8''W).

Figure 8.6.25. Light fixture anchorage in good condition and Gas tubes chained to walls at the
HRZ Hospital, Manta. April 17 th. (GPS coordinates: 0°57'21.8''S, 80°44’308''W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-54
CONCLUSIONS

Non-structural components experienced extensive damage in the area. This damage made
a significant impact on the operations and economy of the cities. Businesses could not operate,
including banks, restaurants, and stores. Hospitals and an Airport terminal buildings were shut
down, relocation operations within the area.

Most of the observed damage has been identified in the FEMA guidelines where suggested
measures are recommended. As the reconnaissance team members observed, and as is the case
in most countries, many of the non-structural components were installed by non-engineers after
construction without consideration of earthquake hazard.

A common misperception is that building utilities and critical systems are heavy enough to
withstand earthquake shaking, but in reality these non-structural components can cause
significant damage from pounding against adjacent objects or falling over, especially under
near-field pulse-like motions of high accelerations as in this Ecuador earthquake. It is critical
to increase public awareness of the risks and hazards of non-structural systems in earthquakes.

The observed damage following the Ecuador earthquake highlights the importance of
seismic design of non-structural components for life safety and post-earthquake operations and
the need to increase educational efforts in this direction. The detailed reconnaissance
information presented in this report for both acceleration- and displacement-sensitive non-
structural components, together with several recorded strong ground motions in the immediate
vicinity of the collected information, presents an excellent opportunity to enhance our
knowledge on the behavior of non-structural components and develop simple, engineering and
common-sense solutions to minimize their seismic risk exposure.

8.7 Residential and Commercial Buildings

INTRODUCTION

Most structures in the areas of the focus reconnaissance locations were low-rise residential
and commercial buildings. Residential properties consisted of reinforced concrete low-rise
buildings as well as one or two-story wood framed houses. Commercial buildings were
typically low-rise or mid-rise reinforced concrete structures. This section focuses on various
types of buildings and the respective damage observed.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-55
TYPICAL RESIDENTIAL HOUSING

Typical residential buildings are 2 to 3-story reinforced concrete structures with an


additional metal deck roof; the first floor is normally commercial. Often, this type of building
is originally built as a 2-story building and a concrete floor is added later on; eventually, a
metal deck roof extension is constructed and the previous roof becomes a terrace. Several of
these residential buildings have 1-3 floors cantilevering more than 5 ft over the sidewalk. If
they do not cantilever, they are supported by slender columns.

Figure 8.7.1. a) Typical residential building with a commercial ground floor in Manta (GPS
coordinates: 0°57’14.4”S, 80°42’36.6”W). b) Residential building in Tarqui, Manta (GPS
coordinates: 0°57’16.83”S, 80°42’32.1”W)

For these typical residential buildings, most of the significant structural damage was soft
story due to the change in stiffness or mass caused by an irregularity or discontinuity in a story.
The reconnaissance team observed collapse of the first or second stories of several buildings.
For numerous buildings that did not have partial soft story floor collapse, there was severe
hinging at the columns. Soft story failure could additionally have been affected by the large
mass from the nonstructural exterior walls.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-56
Figure 8.7.2. a) Typical residential building where floors were added to the original structure.
The third floor seems to have been added after the original construction of the two story
building. b) At the third story, there was severe hinging at the top of the columns (GPS
coordinates: 0°57’20.1”S, 80°42’54.4”W).

Figure 8.7.3. Four-story residential building in Pedernales with commercial first floor. The
second story collapsed due to soft story (GPS coordinates: 0°4’7.3”N, 80°3’18.9”W).

In the areas that are less developed such as in Canoa, there are wood framed 2-story houses
with metal decks roofs. Some of these houses had partial or complete collapse.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-57
Figure 8.7.4. Typical 2-story wood houses in Canoa. Some houses completely collapsed as
shown. (GPS coordinates: 0°27’46.9”S, 80°27’15.0”W)

TYPICAL COMMERCIAL BUILDINGS

Typical commercial buildings are 4 to 7-story reinforced concrete structures comprised of


distinct slab and beams systems; fewer are steel framed. Some commercial buildings suffered
from severe cracking, column hinging, and soft story partial collapse. There was also
significant damage in the nonstructural elements. Specifically, the falling of interior and
exterior nonstructural walls caused major destruction and was life threatening.

Figure 8.7.5. Commercial reinforced concrete building. First, second and fourth floors
collapsed completely. (GPS coordinates: 1°3’20.3”S, 80°27’15.2”W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-58
Figure 8.7.6. Commercial reinforced concrete building in Portoviejo. Several pieces of the
facade collapsed (GPS coordinates: 1°3’18.9”S, 80°27’15.2”W)

Figure 8.7.7. Mall in Portoviejo with a spalled façade (GPS coordinates: 1°3’43.6”S,
80°27’55.3”W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-59
Figure 8.7.8. Partial collapse of nonstructural walls in governmental building in Portoviejo.
This building is occupied by SRI, the Service of Internal Rentals (GPS coordinates:
1°2’60.0”S, 80°27’16.0”W).

CHURCHES

The reconnaissance team visited a church in Pedernales that had significant damage. The
metal roof and the exterior façade had partially collapsed.

Figure 8.7.9. Exterior façade and roof partial collapse in church. The church is located in the
central plaza in Pedernales along the Ruta del Spondylus (GPS coordinates: 0°4’12.4”N,
80°3’15.3”W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-60
8.8 Essential/Critical Buildings

INTRODUCTION

During and after a natural disaster, it is imperative that the essential facilities in the affected
region remain operational. The International Building Code (IBC 2015) defines essential
facilities as buildings and other structures that are intended to remain operational. These
include but are not limited to hospitals, emergency centers, police stations and aviation control
towers. These essential facilities are critical to post-earthquake response and it is crucial that
they remain operational. While not categorized as essential facilities, schools and jails are also
critical structures. These can serve as shelters during post-earthquake recovery.

One way the building codes have for ensuring proper seismic response of these essential
facilities is by requiring the design seismic load to be amplified according to the importance
(risk category) of the building under consideration. For example, the importance assigned to
essential facilities, triggers in most building codes around the world (e.g. ASCE 7, NSR-10)
other requirements associated to the detailing of the structural elements, as well as the design
of the nonstructural elements. The former set of recommendations aims to guarantee the ductile
response of the structural system with low damage, while the latter aims to guarantee the
continuity of the services provided by the facility soon after the event is concluded.

This section focuses on the damage that occurred in hospitals, emergency centers, schools
and airports during the seismic event.

HOSPITALS

Hospitals and clinics are crucial to post-earthquake response and recovery. These should
remain operational after a seismic event to serve injured victims in addition to present patients.
The following map shows the locations of hospitals in the provinces of Manabi and Santo
Domingo de los Tsáchilas, the most affected regions during the earthquake. The
reconnaissance team visited the IESS (Ecuadorian Institute of Social Security) and Rodriguez
Zambrano Hospitals in the city of Manta. These structures are described in detail in Section
8.10.3.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-61
Figure 8.8.1. Hospitals in Manabi and Santo Domingo de los Tsachilas. GEO Salud.
.Ministerio de Salud Publica Ecuador. ©2013-2015 MSP
https://aplicaciones.msp.gob.ec/salud/publico/dniscg/geosalud/gui/#

Figure 8.8.2. Hospitals in Manta City visited by the reconnaissance team. GEO Salud -
Ministerio de Salud Publica Ecuador ©2013-2015 MSP.
https://aplicaciones.msp.gob.ec/salud/publico/dniscg/geosalud/gui/#

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-62
EMERGENCY CENTERS

Emergency centers are crucial in post-earthquake response and recovery, especially


following a natural disaster. The reconnaissance team visited the 911 ECU center and the Unity
of Community Surveillance (Unidad de Vigilencia Comunitaria in Spanish) in Portoviejo.

911 ECU Center

There are several 911 ECU Centers located in distinct cities in Ecuador. These are
multipurpose institutions that focus on emergency response and coordination in the country.
As such, it is imperative that these are operational after any natural disaster.

The 911 ECU Center of Portoviejo is located in the 15 de Abril Avenue. One of its main
functions is to respond to emergency calls and provide necessary assistance. It is a 3-story steel
building structure. This structure was not significantly damaged. From outside it was observed
that a large crack had developed at the roof level in the parapet. Also there were long cracks at
the interior and exterior nonstructural walls in this building. The center was closed for a short
period of time following the earthquake but soon after it was operational.

Figure 8.8.3. The 911 ECU center in Portoviejo. There was a large crack at the roof parapet
(GPS coordinates: 1°4’16.5”S, 80°26’47.5”W).

Unity of Community Surveillance (Unidad de Vigilancia Comunitaria)

The reconnaissance team passed briefly by the Unity of Community Surveillance in


Portoviejo. This 2-year old center accommodates police personnel who are responsible for the
surveillance of the Portoviejo community. Some spalling of the exterior nonstructural walls
was observed from far.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-63
Figure 8.8.4. The Unity of Community Surveillance in Portoviejo. Minor spalling of
nonstructural exterior wall (GPS coordinates: 1°4’16.5”S, 80°26’47.5”W).

SCHOOLS

Schools are crucial facilities that should be occupiable following an earthquake because
they can serve as shelters during periods of response and recovery. Furthermore, in the event
of failure it represents substantial hazard to human life because of its occupant load (IBC 2015).
The reconnaissance team visited four schools that had minor to severe damage.

Unidad Educativa “Linus Pauling”

The Education Unity (Unidad Educativa in Spanish) “Linus Pauling” is a school located in
108 Street and 111 Avenue in Tarqui, Manta. This school was comprised of the original 5-
story reinforced concrete building attached to a newer building of the same height. The older
portion of the building completely collapsed; the newer portion did not have significant visible
structural damage except for the area where it attached to the older building.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-64
Collapsed portion

a)

b)
Figure 8.8.5. a) Unidad Educativa “Linus Pauling” in February 2015 (Google Maps ©2015
Google Inc.). b) Image taken during reconnaissance observation. The older portion of this
structure collapsed (GPS coordinates: 0°57’23.3”S, 80°42’41.0”W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-65
a)

b)
Figure 8.8.6. a) Side view of Unidad Educativa “Linus Pauling” in February 2015 (Google
Maps ©2015 Google Inc.). b) Reconnaissance photos of collapsed building after the earthquake
(GPS coordinates: 0°57’23.3”S, 80°42’41.0”W).

Jardín de Infantes Fiscal Mixto Areli Lopez Pita

This is a nursery school located at Lopez Castillo and Eloy Alfaro in Pedernales. The
reconnaissance team did not enter the establishment to inspect the damages. However, it was
observed that the neighboring residential 4-story building had collapsed and the debris had
already been cleaned out.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-66
a)

b)
Figure 8.8.7. a) Nursery school and neighboring hotel building in Pedernales before the
earthquake. (Google Satellite Views, April 2015 Google Maps ©2015 Google Inc.). b) The
building next to the nursery school collapsed, was demolished and cleaned out. (GPS
coordinates: 0°4’11.7”N, 80°3’20.7”W)

“Unidad Educativa Salesiana San Jose“ School

The “Unidad Educativa Salesiana San Jose” School was opened more than sixty years ago
in Manta. This school educates elementary and high school students. Based on reconnaissance
observations from the exterior of the building, there were large cracks on the exterior walls and
a parapet had fallen along the length of the exterior wall. The reconnaissance team was later
informed that the facility had interior structural damage and some areas were going to be
demolished.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-67
a) b)
Figure 8.8.8. a) Collapsed parapet along the exterior wall. b) Cracked nonstructural walls (GPS
coordinates: 0°57’14.9”S, 80°42’45”W).

Colegio Cristo Rey

This school is located on Cristo Rey Street in the city of Portoviejo. No major dama ge was
observed in this school except for cracks and spalling in some interior and exterior
nonstructural walls.

a) b)
Figure 8.8.9. a) Cracking and minor spalling in concrete nonstructural interior wall. b)
Cracking in masonry nonstructural exterior wall (GPS coordinates: 1°3’16.1S, 80°26’42.2”W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-68
AIRPORTS

After a natural disaster it is imperative that airports and control towers remain operational.
In order to provide the necessary response and recovery resources as well as to transport the
injured to other regions, these transportation facilities should be functional.

The reconnaissance team visited two of the airports in the Manabí area, The Eloy Alfaro
and the Los Perales airports. Both of them were in operation, although the terminal of the Eloy
Alfaro airport was closed soon after the earthquake due to the damages. Flights frequency
increased significantly in both airports, since these were serving flights for emergency and
post-disaster humanitarian assistance after the earthquake.

Figure 8.8.10. Airport locations in the Manabí Area. Google Earth. .©2015.

International Airport Eloy Alfaro, Manta

The Eloy Alfaro Airport is an international airport located on the northeast corner of Manta
City. Eloy Alfaro International Airport is a combination civilian airport and military air base
on the northwest coast area of Manta in the province of Manabí. The airport is more than 35

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-69
years old. It currently operates planes from the armed forces of Ecuador and serves public and
military flights. It is one of the busiest airport in Ecuador. After the earthquake, the airport
terminal closed down. The runway was operational but only served flights with post-disaster
humanitarian assistance.

The airport terminal is a reinforced concrete structure with concrete frames. The slabs are
waffle and the original form blocks have been left. Structurally, some beam to column failure
was observed. The control tower of this airport was a 5-story building that completely collapsed
during the earthquake. There was significant nonstructural damage including fallen ceilings
and partial collapse of nonstructural walls in the Terminal building.

Figure 8.8.11. Eloy Alfaro Airport, Manta. Google Earth.©2015.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-70
Figure 8.8.12. Evacuated International Airport Eloy Alfaro (GPS coordinates: 0°57’10.7”S,
80°41’3.2”W)

Figure 8.8.13. Beam to column connection failure. The rebar at the column is slightly bent and
there is significant spalling present (GPS coordinates: 0°57’11.7”S, 80°41’3.6”W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-71
a) b)
Figure 8.8.14. Cracked and collapsed nonstructural brick walls in the International Airport
Eloy Alfaro (GPS coordinates: 0°57’10.7”S, 80°41’3.2”W)

Figure 8.8.15. Partial collapse of a nonstructural wall. There were dowels from the columns
embedded in the wall. (GPS coordinates: 0°57’10.7”S, 80°41’3.2”W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-72
a)

b)

Figure 8.8.16. a) International Airport Eloy Alfaro Control Tower (Google Satellite View,
February 2015 Google Maps ©2015 Google Inc.) b) The location of the control tower which
has been cleaned out after collapse (GPS coordinates: 0°57’9.5”S, 80°41’1.9 ”W).

Los Perales Airport, San Vicente

Los Perales Airport was constructed in 1950 in San Vicente, near Bahía de Caráquez city.
The original runway was approximately 500 meters long. Years later during Duran’s
presidency, a 1100 meter long extension was added to the runway. In the past few years, this
airport has been a controversial topic since it was in the verge of being closed permanently.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-73
However, the people of the area wanted the airport to remain operational due to its proximity
to the region. This airport has not served commercial flights since 2000 (El Diario News
http://www.eldiario.ec/noticias-manabi-ecuador/261715-pista-de-los-perales-funciona-a-
medias/). This airport was operational after the earthquake. After the earthquake, the airport
served more than hundred flights a day.

The reconnaissance team visited this airport and observed minor structural damage. In the
airport terminal, pounding of two adjacent sections of the structure caused collapse of the
parapet, which fell on a car during the earthquake. It was also observed that the extended
portion of the runway had cracks running along the runway. (El Comercio News
http://www.elcomercio.com/actualidad/aeropuertos-manta-bahia-operan-restricciones.html).

Figure 8.8.17. Los Perales Airport, San Vicente. Google Earth ©2015.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-74
Figure 8.8.18. Pounding and collapse of parapet at the Airport of the Perales (GPS coordinates:
0°36’5.3”S, 80°24’27.0”W).

Figure 8.8.19. Pounding of columns and roof parapet at the Airport of the Perales terminal
(GPS coordinates: 0°36’5.3”S, 80°24’27.0”W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-75
Figure 8.8.20. The newer portion of the runway at the Airport of the Perales had cracks along
the runway (GPS coordinates: 0°36’58.7”S, 80°23’53.8”W).

8.9 Special Structures, Tanks, Camaroneras


This Chapter is not included in the Rapid Report. It will be developed in the Final Ecuador
GEER Reconnaissance Report.

8.10 Case Studies

8.10.1 Case Studies: Bridges


THE BAHIA-SAN VICENTE, LOS CARAS BRIDGE

Figure 8.10.1.1. Bridge site looking eastward from San Vicente to Bahía de Caráquez (GPS
coordinates: 0°36'30.66"S ,80°24'30.59"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-76
Figure 8.10.1.1 shows the important bridge that joins the coastal towns of San Vicente and
Bahía de Caráquez, which is located in Manabí province and facilitates the transportation. The
project was undertaken by Ecuador's Army Corps of Engineers and was built in 2010. The
bridge consists of a central span from abutment one to abutment two for a total length of
1,984.5m. The entry to bridge is via a 208m access ramp from the Bahía span, and a 607m
access ramp from the San Vicente span. The bridge is supported by 48 piers and two abutments
divided in three sections (Figure 8.10.1.2):

 Bahía de Caráquez Access: Abutment E1 and Piers P-1 through P-6, a total of 120.3
m length and 10.2 m wide.
 Central Section: Piers P-6 thru P-44, a total of 1,710 meters (38 piers), the pier
separation at the central sections is 45 meters. There are two kinds of continuous
lengths, four continuous spans, with 180 m length, and two continuous spans, with 90
m length, 13.2 m wide

 San Vicente Access: Piers P-44 through P-48 and abutment E-2, a total of 153.95 m
length and 13.2 m wide.

Figure 8.10.1.2. Plan of view of Los Caras Bridge, Courtesy of Ecuador’s Army Corps of
Engineering (GPS coordinates: 0°36'33.58"S, 80°24'58.68"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-77
The superstructure of the bridge was designed on the basis of pre-stressed pre-tensioned
and post-tensioned concrete. The bridge deck is 9.3m wide for vehicular traffic and has a 3m
pedestrian and bicycle lane. The roadway section has two traffic lanes. Each of the bridge’s
concrete slab spans consist of six beams, and of these, two are directly supported by seismic
isolators (152 Triple Pendulum in the central section). The bridge site is characterized as an
estuarine environment formed by alluvial deposits with depths between 25 and 85 meters
overlying weathered rocks. The design team decided to build the bridge foundations with open-
end steel pipe piles, mainly supported by skin friction. The bridge cross section is illustrated in
the figure 8.10.1.3.

Figure 8.10.1.3. Bridge cross-section (Courtesy of Ecuador’s Army Corps of Engineers)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-78
The foundations for the central section were built with open-end steel pipe piles, 1.21
meters in diameter and 20 mm wall thickness, with variable mean lengths ranging from 32 to
65 meters. Figure 8.10.1.4 illustrate the soil profile of Los Caras Bridge.

Figure 8.10.1.4. Soil profile Los Caras Bridge (Courtesy of Ecuador’s Army Corps of
Engineers and Adolfo Caicedo)

The bridge is isolated with the use of Triple Friction Pendulum Seismic Isolation which
exhibit significant non-linear behavior. The devices are designed to prevent the transfer of large
seismic forces to the superstructure by decoupling it from the substructure. Therefore, there is
reduction of forces (accelerations) in the superstructure and substructure, and force
redistribution between the piers and the abutments. A seismic excitation would cause the
maximum displacement of 58.4 cm determined by the isolator’s Lateral Displacement
Capacity.

Observations made after the earthquake:

Some of these comments were made by the engineer Marcelo Romo, responsible for the
design as part of the Army Corps of Engineers, after a visual inspection of the bridge and its
components).

The bridge was operable and there was no evidence of damage requiring closure of the
bridge. In figure 8.10.1.5, the technicians from the Army Corps of Engineers show the
evaluation carried out by the Ministry of Urban Development and Housing of Ecuador
(MIDUVI). The Army Corps of Engineers presented a document of the inspection made by the
MIDUVI.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-79
Figure 8.10.1.5 Ecuador’s Army Corps of Engineers team shows the document inspection of
MIDUVI (GPS coordinates: 0°36'48.85"S, 80°24'46.93"W)

There was no visible damage to the modular long expansion joint at the west (Bahía de
Caráquez) end of the bridge (Figure 8.10.1.6). At the east (San Vicente) end, there was a
detachment of a neoprene section and the expansion of the joints (Figure 8.10.1.7). Near the
middle of the bridge, there was no evidence of damage (Figure 8.10.1.8). There was no
significant cracking of the concrete deck which might have indicated that the piers had settled.

Figure 8.10.1.6. Modular bridge long expansion joint (GPS coordinates: 0°36'48.85"S,
80°24'46.93"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-80
Figure 8.10.1.7. Modular bridge long expansion joint, San Vincente side (GPS coordinates:
0°36'30.74"S, 80°24'30.54"W)

Figure 8.10.1.8. Neoprene Rubber Standard Expansion joint (GPS coordinates: 0°36'35.28"S,
80°25'17.15"W)

Figure 8.10.1.9 illustrate the impact absorbers or external movement synchronizers that
reflect at least 5 important impacts during the quake.

Figure 8.10.1.9. External movements synchronizers (GPS coordinates: 0°36'33.58"S,


80°24'58.68"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-81
The elastomeric bearings of Bahía and San Vicente approaches did not show evidence of
damage. There are cracks on the piers 3, 4 and 5 (from the south end of the Bahía de Caráquez
access) that need to be repaired. The horizontal shear blocks were cracked while retraining the
lateral movement of the superstructure, Figure 10.8.10.

Figure 8.10.1.10. Cracking of the pier 3 on the horizontal shear block at the Bahia side, the pier has
elastomeric bearings (GPS coordinates: 0°36'36.56"S and 80°25'27.26"W)

All piers, but pier 12 and its neighbors had a relatively uniform behavior in the earthquake.
The bearings indicate that a lateral displacement took place. The lateral bearing toward the side
face of the bridge appear to show a larger travel from South to North, about 50% higher. Pier
12 was the only pier that showed much larger displacements. The base isolation system did not
have any damage. According to the inspection carried out by the team of the Ecuadorian Army
Corps of Engineers. the average differences were in the range of 15 cm in the north direction
and 30 cm in the southern direction (Figure 8.10.1.11).

Figure 8.10.1.11. Friction pendulum device as part of the base isolation system of Los Caras Bridge
(Left), Inner displacement isolator view (GPS coordinates: 0°36'33.58"S, 80°24'58.68"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-82
8.10.2 Case Studies: Ports
This Chapter is not included in the Rapid Report. It will be developed in the Final Ecuador
GEER Reconnaissance Report.

8.10.3 Case Studies: Hospitals, Schools/Critical Facilities


INTRODUCTION

Throughout history, Ecuador has suffered devastating economic, social, and environmental
consequences from natural disasters. The vulnerability of Ecuador is apparent from numerous
disastrous events that have included earthquakes, floods, landslides, mudflows, hurricanes and
volcanic eruptions. Undoubtedly, one of the worst impacts resulting from natural disasters
is the effect on health infrastructure, principally hospitals. For example, Ecuador suffered
major earthquakes in 1987 and 1998 that severely damaged hospitals such as José Maria
Velasco Ibarra and Miguel H. Alcívar hospitals. These hospitals were not adequately prepared
for such events. The use of modern building codes can limit damage to structures but non-
structural elements and essential equipment in buildings remain at risk. In critical structures,
especially hospitals, sub-par performance of non-structural systems can lead to consequences
as devastating as structural failures. Noting that non-structural elements in the hospitals
suffered major damage, therefore these structures could not continue to provide services during
times of crisis. The hospital has critical medical equipment and complex mechanical and
electrical systems which must have guaranteed functionality because the hospital cannot
operate without them.

IESS HOSPITAL, MANTA

The IESS is the Ecuadorian Institute of Social Security that has numerous hospitals
throughout Ecuador. The IESS Hospital in Manta is located at the Avenida de la Cultura. This
hospital was severely affected by the earthquake causing a total of five patient deaths. There
were both significant structural and nonstructural damages that contributed to the complete
evacuation of this hospital. Supply of electricity was not guaranteed during the earthquake,
which was associated to some of the casualties in the intensive care unit (ICU).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-83
Figure 8.10.3.1. IESS Hospital Location, aerial view, Google Earth 2016. (GPS coordinates:
0°57'17.90"S, 80°43'26.70"W).
The hospital building is divided into two sections; a 2-story and a 5-story reinforced
concrete building. The 2-story building has a metal deck roof and a clerestory with skylights
on distinct portions.

Figure 8.10.3.2. Front façade of the IESS Hospital in Manta (GPS coordinates: 0°57'17.90"S,
80°43'26.70"W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-84
Construction Means and Methods

Albeit being considered a non-informal construction, the observed quality of erection is


considered low. Evidence of rebar exposure due to lack of proper concrete cover was observed,
along with the presence of cold joints along the length of the beams and and/or near beam-
column joints.

Figure 8.10.3.3. Rebar exposure and cold joint presence in beams of the IESS Hospital in
Manta (GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

Structural Damages

In the 2-story portion of the building, there was significant formation of hinges in the beam-
column joint at the top of the 1st floor columns. In the 5-story portion there was spalling of
concrete in some beams and stair slabs. Some structural damage was observed at the beam-
column joint where diagonal cracks are evidence of large shear demand. Additionally, some
beams in the third story showed signs of large flexural cracks at the beam-column interface.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-85
Figure 8.10.3.4. Damage above columns on the 2-story portion of the IESS Hospital in Manta
(GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

Figure 8.10.3.5. Cracks in or near the beam-column joints at the 2 nd story of the 5-story portion
of the IESS Hospital in Manta (GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

Located near the mid span of the building are the stairs. These occupy the right side of the
bay, causing an unsymmetrical distribution of stiffness which attracts more shear force into the
right side of the frame.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-86
Figure 8.10.3.6. Staircase state from the 2 nd to 5th story of the 5-story portion of the IESS
Hospital in Manta (GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

Non-structural Damages

Most of the damage in the hospital was caused by the failure of nonstructural components.
The main damages were contributed by the spalling and falling of the brick walls. Also, there
were many cases of falling ceilings, ducts and pipes.

- Non-structural Walls
The interior and exterior walls were mainly composed of hollow clay and concrete block.
Overall there was a combination of out-of-plane and in-plane failure. Several exterior walls
and windows and some interior walls completely fell over. Most interior walls were free
standing; the walls were not full height and were not braced in any horizontal direction. The
interior walls had significant cross tension cracks and spalling. The reversal in the shear force
at the walls during the earthquake caused the diagonal cracks in both directions. See Section
11.6 for more information regarding nonstructural damage.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-87
Figure 8.10.3.7. Appearance of concrete bricks used in the façade of the IESS Hospital in
Manta (GPS coordinates: 0°57'17.90"S, 80°43'26.70"W).

Figure 8.10.3.8. Cross tension cracks and spalling in hollow clay interior wall in the IESS
Hospital in Manta (GPS coordinates: 0°57'17.3"S, 80°43'25.2"W).

Figure 8.10.3.9. a) Collapse of non-structural concrete block exterior walls in the IESS
Hospital in Manta. The interior nonstructural walls are clearly not full-height and are not braced
horizontally (GPS coordinates: 0°57'18.1"S, 80°43'25.3"W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-88
- Ceilings, Ducts and Pipes

Many of the ceilings, duct and pipes partially or completely fell, causing life threatening
conditions. No seismic horizontal bracing was observed.

Figure 8.10.3.10. Fallen parts of the ceiling in the IESS Hospital in Manta (GPS coordinates:
0°57'18.1"S, 80°43'25.3"W).

Figure 8.10.3.11. Fallen ducts and ceilings in the IESS Hospital in Manta (GPS coordinates:
0°57'18.1"S, 80°43'25.3"W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-89
RODRÍGUEZ ZAMBRANO HOSPITAL, MANTA

The Rodríguez Zambrano Hospital was founded in 1911 in Manta. It is located in the Buena
Vista neighborhood. This hospital in Manta had minor to severe damages. Initially most of the
hospital was evacuated. Currently, it is opened although it is not operational.

The hospital structure is divided into a few reinforced concrete structures ranging from a
3-story building to a 7-story building. Part of this hospital was constructed with moment
resisting concrete frames with waffle slabs. The 3-story building had a steel frame addition
with a metal deck roof.

Figure 8.10.3.12.. Front façade of the Rodríguez Zambrano Hospital in Manta (GPS
coordinates: 0°57'14.2"S, 80°4330.9"W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-90
Figure 8.10.3.13.. Reinforced concrete construction with moment resisting concrete frames
with waffle slabs in the Rodríguez Zambrano Hospital in Manta (GPS coordinates:
0°57'21.8"S, 80°44'30.8"W).

Structural Damages

Overall, the structural skeleton of this building observed by the reconnaissance team was
in good condition.

Nonstructural Damages

Some interior nonstructural walls had diagonal and cross tension cracks but most did not
have significant spalling. On the 6th floor, there were some interior walls that fell. During the
earthquake, some of ceilings remained in place.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-91
Figure 8.10.3.14. Cross tension cracks in interior nonstructural walls in the Rodríguez
Zambrano Hospital in Manta (GPS coordinates: 0°57'15.8"S, 80°44'32.6"W).

Figure 8.10.3.15.. Spalling in hollow clay interior walls in the Rodríguez Zambrano Hospital
in Manta (GPS coordinates: 0°57'19.3"S, 80°44'31.53"W).

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-92
Figure 8.10.3.16. The highlighted area used to have nonstructural walls that collapsed during
the earthquake. This is the 6th floor in the Rodríguez Zambrano Hospital in Manta. Refer to
Section 11.6 (GPS coordinates: 0°57'14.9”S, 80°44'31.9"W).

A large lighting equipment was well anchored into the ceiling and did not fall (See Figure
8.10.3.17). Some tanks were chained to the wall which did not allow them to tip over. Many
of the shelves with records and files had fallen to the floor. When the reconnaissance team
visited the hospital, the hospital staff had already started to effectively attach these
nonstructural components to the walls so that they wouldn’t fall due to aftershocks or during
the next large earthquake. Refer to section 11.6.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-93
a) b)
Figure 8.10.3.17. Medical equipment in the Rodríguez Zambrano Hospital in Manta. a) A large
lighting equipment was well anchored into the ceiling and did not fall during the earthquake.
b) Tanks in this hospital room were chained to the wall so that they would not fall (GPS
coordinates: 0°57'21.8"S, 80°44'30.7"W).

MIGUEL HILARIO ALCÍVAR HOSPITAL, BAHÍA DE CARÁQUEZ

Miguel H. Alcívar Hospital is located in the city of Bahia de Caráquez with a capacity of
120 beds. It is noteworthy that the hospital in 1998 earthquake suffered considerable damage
to its structural and non-structural components, making it inoperable for a period of four years
(See figure 8.10.3.18). After the 1998 earthquake, the Ecuador Ministry of Public
Health undertook a retrofit project that consisted of creating a monolithic structure with the
existing blocks, based on jacketed columns and concrete shear walls, with the intent of
strengthening the structure.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-94
Figure 8.10.3.18. Structural and non-structural damage Miguel H. Alcívar hospital during
the1998 earthquake in Ecuador (Courtesy of Aguiar R.).

The Hospital is located in the province of Manabí, in the urban center of Canton Sucre (See
Figure 8.10.3.19.)

Figure 8.10.3.19. Miguel H. Alcívar Hospital (GPS coordinates: 0°37'18.37"S,


80°25'39.52"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-95
This hospital was affected by the 2016 earthquake causing a total of one death. This hospital
suffered no structural damage, the hospital was completely closed because there was an
extensive nonstructural damage related to nonstructural elements that are sensitive to drifts.
The most common structural damage was due to fallen partition walls at the ground and first
floor, objects that were anchored. The level of performance of non-structural elements and
components was inadequate (See figures 8.10.3.20 to 8.10.3.22), disabling the hospitals and
putting occupants at risk during and after April 16th earthquake.

Figure 8.10.3.20. Non-structural damage: Masonry infill - Outpatient (GPS coordinates:


0°37'18.46"S, 80°25'38.62"W)

Figure 8.10.3.21. Non-structural damage: Masonry infill- Outpatient (GPS coordinates:


0°37'18.20"S, 80°25'38.95"W)

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-96
Figure 8.10.3.22. Non-structural damage contents: Ground level (GPS coordinates:
0°37'18.05"S, 80°25'39.00"W)

8.10.4 Case Studies: Dams


This Chapter is not included in the Rapid Report. It will be developed in the Final Ecuador
GEER Reconnaissance Report.

8.10.5 Case Studies: Seismically Isolated Buildings


This Chapter is not included in the Rapid Report. It will be developed in the Final Ecuador
GEER Reconnaissance Report.

8.10.6 Case Studies: Cementeries


This Chapter is not included in the Rapid Report. It will be developed in the Final Ecuador
GEER Reconnaissance Report.

GEER-ATC Muisne, Ecuador, Earthquake 8 Structural Observations


Report Version 8-97
Blank Page
CHAPTER 9
Community Preparedness
and Response

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
9.1 Introduction and the Red Cross
The information presented in this chapter is based on data provided by the National Risk
Management Agency (Secretaría de Gestión de Riesgo - SGR) of Ecuador. The immediate
response, action, and community involvement following the main M w7.8 event of April 16 th
are presented for a number of provinces and cities including the provinces of Esmeraldas and
Manabi which were the most affected (Fig. 9.1.1).

Figure 9.1.1. Map depicting the main cities of Ecuador (photo from web, gosouthamerica.com).

Historically, the Ecuadorian Red Cross (Cruz Roja Ecuatoriana), CRE, has played an active
role in the risk management of emergencies. As an auxiliary entity to public authorities, the
CRE has always been active when natural or manmade disasters have occurred. The procedure
CRE follows focuses on building strong communities and reducing future vulnerabilities by
developing preparedness plans, community risk maps, and contingency plans for better and
more efficient future disaster responses. To achieve their goals, the CRE receives periodic
reports from the Communications Centre (Radio de Telecomunicaciones) which handles
simultaneous information from 32 monitoring stations across Ecuador.

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-1
9.2 Army Action Plan and Immediate Response
THE DAY OF THE EARTHQUAKE

Immediately after the April 16 th earthquake, the government declared a “State of


Exception” over the national territory, with the six provinces of Esmeraldas, Santo Domingo,
Manabí, Guayas, Los Ríos, and Santa Elena placed in a “State of Emergency” which required
mobilizing security forces to those provinces. The healthcare system was placed on high alert
to treat the injured. School activities were canceled in the six provinces, and a national soccer
tournament, a very popular event in the country, was canceled (Dialogo Americas, 2016).
International aid was requested by the government.

The Emergency Operations Committee (Comité de Operaciones de Emergencia), COE,


activated all their taskforces and corresponding contingency plans and states of alert for
hospitals. An initial group of 50 rescue and recovery workers with 25 volunteers was
dispatched throughout the affected provinces to survey damage (Fig. 9.2.1). The immediate
response in the main cities on the day of the main event was as follows:

 Quito (capital), Pichincha province: The hospitals Pablo Arturo Suarez, Eugenio Espejo,
Nueva Aurora, and the Elderly (Hospital del Adulto Mayor) were placed in high alert.
 Esmeraldas province: The whole coastal area was evacuated as a precaution.
 Manabí province: Evacuation was performed as a precaution to higher areas.
 Balzar County, Guayas province: People were transferred to assistance homes.
 Chimborazo province: The hospital was partially functional, with fissures in walls, glass.

Figure 9.2.1. Emergency personnel and volunteers trying to save a person trapped in car in Guayaquil
right after the earthquake on April 16th. Photo: R. Cedeño (Vistazo Magazine, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-2
FIRST DAYS AFTER THE EARTHQUAKE

On April 17th, 30 ambulances from the Ministry of Public Health MSP (Ministerio de Salud
Pública), the Ecuadorian Social Security Institute IESS (Instituto Ecuatoriano de Seguridad
Social), and the CRE were dispatched to affected areas. Volunteers helped vigorously in search
of survivors (Fig. 9.2.2).

Figure 9.2.2. Volunteers search for survivors in the debris of buildings in Pedernales on April 17th.
Photo by US Department of State (blogs.state.gov/stories/2016/04/17/earthquake-ecuador).

The Armed Forces joined with 4 helicopters (Fig. 9.2.3) and a group of 10,000 to assist
victims (Dialogo Americas, 2016). The National Police had 2 helicopters available for the
rescuing efforts. Two mobile hospitals were also available with a first response team of 30
people that were joined by a group of volunteers.

Figure 9.2.3. Army personnel deployed with helicopters to affected areas on April 17th, 2016. Photo:
Ecuadorian Ministry of Defense (Dialogo Americas, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-3
The National Secretariat of Planning and Development (Secretaría Nacional de
Planificacion y Desarrollo) divides Ecuador in nine zones assigned to Rapid Response Teams
after a disaster (SENPLADES, 2012). Zones 1, 4, 5, and 8 on Fig. 9.2.4 were activated. The
actions taken are listed below with details on Tables 9.2.1 and 9.2.2.

 Manabí province: Teams from MIES, National Police, Firefighters, MSP, and Armed
forces met at Manta to distribute aiding groups in the province.
 Ambato, Tungurahua province: One Mobile hospital was sent to Pedernales.
 Guayaquil, Guayas province: One Mobile hospital was sent to Portoviejo.

Figure 9.2.4. Zones assigned to Rapid Response Teams (SENPLADES, 2012).

The CRE remained in Manabí and Esmeraldas and was involved in prehospital care,
sanitation assistance, psychosocial support, restoring family connections, damage assessment,
and evaluation of the needs of the victims (Fig. 9.2.5). A mobile unit was available for medical
assistance, covering the areas of Portoviejo, San Vicente, Jama, and Canoa. There were 39
permanent and 62 temporary shelters available, accommodating 6,804 families and a total of
29,067 people. Search and rescue operations (Fig. 7.2.6) continued for days after the main
event as shown on Figs. 9.2.7, 9.2.8 provided to GEER-ATC by the Ecuadorian Armed Forces.

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-4
Figure 9.2.5. CRE volunteer providing pre-hospital care and implementing family links restoration
program available for people affected by the earthquake. Photo by CRE.

Table 9.2.1. Emergency response personnel deployed one day after the earthquake (SGR-14, 2016).
Referenced agencies: (*) EDAN: Damage Assessment & Needs Analysis (Evaluación de Daños y
Análisis de Necesidades) and (**) SSC: Rural Social Security (Seguro Social Campesino).
# Institution Origin Destination

80 Quito's Firefighters Unit Quito Manta

30 Machala's Firefighters Unit Machala Manta

29 Cuenca's Firefighters Unit Cuenca Manta

13 Santo Domingo's Firefighters Unit Santo Domingo Pedernales

4385 National Police Quito Manta

10000 Armed Forces

37 GIR National Police Quito Manta

150 Police Officers Applicants Quito Manta

114 Red Cross Volunteers Different provinces Manta

20 MSP personnel from Zone 9 Quito Manta

12 Ambulances

12 2 Structural Evaluators Teams

15 2 EDAN* Teams

8 2 MSP Mobile Surgery Units


Ambato Pedernales
96 2 Mobile Hospitals
Guayaquil Portoviejo
30 1 General Central Unit Chimborazo Manta
Portoviejo
60 Doctors from SSC**
Manta
11 Red Cross Ambulances
*EDAN: Damage Assessment and Needs Analysis (Evaluación de Daños y Análisis de
GEER-ATC Muisne, Ecuador, Earthquake Necesidades) 9-Community Preparedness and Response
Report Version 1**SSC: Rural Social Security (Seguro Social Campesino) 9-5
Volunteers
Table 9.2.2. Emergency response Activateddeployed
volunteers for the Emergency
one day after
after the
the Earthquake
earthquake (SGR-14, 2016).
in April 16th 2016
Zone Province Numbers Actions

1 Esmeraldas 4 Assistance in the territory

2 Orellana 5 In alert

Chimborazo 40 In alert

3 Cotopaxi 21 In alert

Tungurahua 18 In alert

Santo Domingo 6 Assistance in the territory


4
Chone 30 Assistance in the territory

Guayas 30 Assistance in the territory


5
Los Ríos 30 Assistance in the territory

Azuay 25 Supporting the governorate


6
Cañar 8 In alert

7 El Oro 41 In alert

Total 258

Figure 9.2.6. A woman and her daughter are rescued Sunday after they were trapped in the rubble from
the earthquake. Photo by National Police of Ecuador (NBC, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-6
Figure 9.2.7. Moment of silence observed by rescue and recovery team in Pedernales (4/19/16, ~GPS
0o5'0''N, 80o7'0''W). Photo provided to GEER-ATC team by the Ecuadorian Armed Forces.

Figure 9.2.8. Army rescue and recovery team in Pedernales (4/19/16, ~GPS 0o5'0''N, 80o7'0''W). Photo
provided to GEER-ATC team by the Ecuadorian Armed Forces.

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-7
9.3 Zone Zero
THE DAY OF THE EARTHQUAKE

 Esmeraldas province: Every hospital was activated. Healthcare Centers were limited in
their operations due to the sudden increasing demand.
 Pedernales, Manabí province: The whole city was declared a Disaster Area.

ONE WEEK AND ONE MONTH AFTER THE EARTHQUAKE

Precautions against disease outbreaks were taken with medical teams actively monitoring
the shelters. In the province of Manabí, the following actions were taken:
 1,511 volunteers remained in Pedernales, Manta, and Portoviejo (Fig. 9.3.1).
 5 mobile hospitals and 111 operative ambulatory units were available in Manabí.
 Mobile hospitals were stations in Bahía de Caráquez (1), Portoviejo (2), and Chone (2).
 2 mobile surgery units were active in Manta and 33 operative units were open in Portoviejo.
 Operative units were working in Portoviejo (33) and Pedernales (4).
 Health care centers Ángeles de Colón, Nuevo Portoviejo, Primero de Mayo, Km90 closed.

Figure 9.3.1. Army cleanup operations in Carapoto, Manabí (4/21/16, ~GPS 0o5'0''S, 80o29'0''W).
Photo provided to GEER-ATC team by the Ecuadorian Armed Forces.

A month after the seismic event, the commercial activity in the area was restored by 60%.

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-8
9.4 Homeless Relocation
Immediately after the earthquake, evacuations were conducted in the most affected areas.
Specifically: (i) the entire coastal area of Esmeraldas was evacuated as a precaution; (ii) the
higher areas of Manabí were evacuated as a precaution to higher areas; (iii) most of the
population of Balzar County in Guayas was transferred to assistance homes.

ONE DAY AFTER THE EARTHQUAKE

Shelters shown on Table 9.4.1 were ready to temporarily accommodate the evacuated
population. Additionally, civilians in Quito provided temporary shelter to few families.

Table 9.4.1. Shelters activated immediately after the earthquake (SGR Status Report 14, 2016).
Shelters Activated In Response to the
Emergency of April 16th 2016
Province Shelters
Esmeraldas 3
Los Ríos 3
Guayas 2
Santo Domingo 1
Manabí 2
Pichincha 1
Total 12

The Ministry of Economic and Social Inclusion (Ministerio de Inclusión Económica y


Social), MIES, organized teams to offer immediate support and coordinated the distribution of
kits and temporary tents (Fig. 9.4.1) as described in Table 9.4.2.

Figure 9.4.1. UNHCR (United Nations Refugee Agency) staff along with police and local people work
install tents in Chamanga, Esmeraldas (~GPS 0o16'0''N, 79o56'0''W). Photo UNHCR/Santiago Arcos.

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-9
Table 9.4.2. MIES emergency response teams (left) and tents distributed (right) after SGR#14 (2016).
MIES National and District Teams in Response to Tents Sent to Affected Areas in Response to
the Emergency of April 16th 2016 the Emergency from April 16th 2016
Province City Destination Teams Province Tents Tent's Capacity Destination
Esmeraldas Imbabura 1 100 Pedernales
Pichincha Quito 8
Pedernales Santo Domingo 1 100 Portoviejo
Guayas Guayaquil Portoviejo 3 Cotopaxi 1 100 Pedernales
Azuay Cuenca Portoviejo 2
Total 13

ONE WEEK AND ONE MONTH AFTER THE EARTHQUAKE

A week after the earthquake, 103 shelters were active in the provinces of Santa Elena,
Bolivar, Pichincha, Manabí, Los Ríos, Imbabura, and Esmeraldas (Table 9.4.3). The shelters
accommodating victims a month after the earthquake are listed in Table 9.4.4.
Table 9.4.3. Active shelters a week after the earthquake (SGR Status Report 41, 2016).

Shelters Activated in Response to the Emergency from April 16th 2016

# Provinces County Families People


1 Santa Elena Santa Elena 7 30
2 Bolívar Guaranda 1 9
3 Santo Domingo La Concordia 158 632
4 Santo Domingo Santo Domingo 61 253
5 Pichincha Quito 46 236
6 Manabí Portoviejo 434 1398
7 Manabí Bahía de Caráquez 126 599
8 Manabí Manta 10 40
9 Manabí Canoa 500 2500
10 Manabí San Vicente 82 350
11 Manabí Pedernales 1368 5955
12 Manabí Jama 963 3789
13 Manabí El Carmen 84 384
14 Manabí Chone 1995 8379
15 Manabí Flavio Alfaro 31 139
16 Manabí Rocafuerte 111 434
17 Manabí Sucre 32 150
18 Los Ríos Babahoyo 64 240
19 Esmeraldas Muisne 708 3458
20 Esmeraldas Esmeraldas 21 85
21 Imbabura Cotacachi 2 7
Total 6804 29067
Total Shelters 103

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-10
Table 9.4.4. Active shelters a month after the earthquake (SGR Status Report 71, 2016).

Families People No. of Shelters

7,319 28,775 251

9.5 Supplies and Medical Services


The Emergency Operations Committee (Comité de Operaciones de Emergencia), COE,
was activated immediately to assist with victims. In order to provide medical treatment, mobile
units were installed outside the premises of hospitals that were damaged and not functional
after the earthquake. An example is the Rafael Rodriguez Zambrano hospital that had to
perform partial evacuation and transfer patients to other facilities (Fig. 9.5.1).

Figure 9.5.1. Treating victims outdoors in Rafael Rodriguez Zambrano Hospital (Teleamazonas, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-11
ONE DAY AFTER THE EARTHQUAKE

Humanitarian assistance was provided by government agencies, such as MIES that worked
with the Worldwide Food Program, PMA, to deliver a food voucher per family which provided
food for 3 days with participation of 5 supermarkets in Manta, 3 in Portoviejo, 1 in Chone, and
1 in Esmeraldas. The water bottling company Tesalia sent 5,000 gallons to Portoviejo, 3,000
gallons to Pedernales, and 3,000 gallons to Esmeraldas (Fig. 9.5.2). Tables 9.5.1 to 9.5.3
provide information on the distributed kits.

Figure 9.5.2. Volunteers and MIES members distribute food supplies (top). Members of the National
Police organize water bottles in Pedernales on April 20 th (bottom). Photo by M. Ayala, (Andes, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-12
Table 9.5.1. Kits sent to Manabi in response to the Emergency (SGR Status Report 14, 2016).

Kits Sent to Manabi in Response to the Emergency


from April 16th 2016

# Kits Packaged kits


1000 Food
678 Sleeping kits
152 Personal hygiene kits
157 Kits of family dishes
200 Clothing kits

Table 9.5.2. Kits sent to Esmeraldas in response to the Emergency (SGR Status Report 14, 2016).

Kits Sent to Esmeraldas in Response to the Emergency


from April 16th 2016

# Kits Packaged kits


584 Food
448 Sleeping kits
158 Personal hygiene kits
156 Kits of family dishes

Table 9.5.3. Kits sent to affected areas in response to the Emergency (SGR Status Report 14, 2016).
Kit Portoviejo Pedernales Esmeraldas
Family Dishes 700 1000
Sleeping 3000 4000

Food 3000 vouchers 3000 vouchers 1000 vouchers

Personal Hygiene 1000 1000


Family Cleaning 250 250

Communal Kitchen Utensils 200 300

Cleaning Temporary
100 100
Accommodation

Clothing 100 100

2980
Mattresses
(Petroecuador)

2980
Mosquito Nets
(Petroecuador)
Plastic
3 (UNICEF) 7 (UNICEF)
Tarps
Tents 4(MCDS) 4 (MCDS)

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-13
ONE WEEK AND ONE MONTH AFTER THE EARTHQUAKE

The CRE established a water plant in a Technical High School at Pedernales that operates
as shelter. They daily distributed 35,000 litres of water with ARCA’s support. At the same
time, the setups of two more water plants were coordinated with the Ecuadorian Red Cross and
Colombian Red Cross. Both had a daily capacity of 45,000 litres (Fig. 9.5.3).

Figure 9.5.3. Potable water distribution on April 20, 2016. Photo by the Ecuadorian Armed Forces.

Figure 9.5.4. Belen Carillo, a UNICEF child protection specialist, with Kimberly, 6, at the “Y” shelter
that hosted 250 displaced families outside of San Jose de Chamanga. Web photo UNICEFUSA/Sandler.

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-14
A fundraising campaign was initiated by the CRE for the acquisition of kits (hygiene, food,
cleaning, medical supplies). In collaboration with the Guayaquil city, they received donations
including 1,920 boxes of water bottles, 945 with food, 945 with hygiene kits, and 1,515 with
clothes. International support arrived, with UNICEF sending an airlift of 10,000 fleece
blankets, 300 plastic tarps, over 100 large tents, 4,000 insecticide treated bed nets, 250,000
Vitamin-A capsules, and diarrhea treatment medication. In the shelters for displaced families,
UNICEF applied the “Back to Happiness program” (Retorno a la Alegria) that provides
psychological support and uses art to help children who have been traumatized by the disaster
regain a sense of normalcy (Figs. 9.5.4 and 9.5.5).

Figure 9.5.5. A young boy, potentially at risk for post-disaster trauma, colors in a child-friendly space
in Portoviejo. Photo 4/19 UNICEF/ECU/Castellanos (unicefusa.org/ecuador-earthquake).

During the first month after the main event, MIES and the Ecuadorian Armed Forces
distributed numerous food kits as presented on Table 9.5.4. Private and educational institutions
and civilians donated supplies and their time to support the people (Fig. 9.5.6).

Table 9.5.4. Food kits from 4/17 to 5/17 provided by MIES and the Armed Forces.

Food Distribution a Month after the Earthquake (4/17-5/17/16)

Week Time Period No. of Kits Provider


4/17  4/20 98,494 MIES
1
4/24 166,030
2 5/1 199,048
Ecuadorian Armed
3 5/8 132,478
Forces
4 5/15 111,662
5 5/17 30,075

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-15
Figure 9.5.6. Pedernales, April 24: Kiara Farias, 2, whose arm was broken in the earthquake plays with
her brother Jostin, 6, in a makeshift camp for displaced people (top). Volunteers serve food to children
(bottom). Photos by R. Abd, AP (web: www.fresnobee.com/news/local/article74243762.html).

9.6 Security
ONE DAY AFTER THE EARTHQUAKE
Following initial damage assessment, precautions were taken to ensure the safety of the
civil population. The instructions given by the National Transit Agency of Ecuador (Agencia
Nacional de Tránsito del Ecuador), ANT, included:

 The National Police (Policía Nacional), PN, forbade entrance to the affected areas and
was responsible for:
Security and control of the population.
Assistance in rescuing the victims.
Control operations taken as a precaution.
Permanent patrol to avoid robbery and looting.
Support in the evacuation for civilians that remained in an unsafe location.
Monitoring of roads and highways near the affected areas.
GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response
Report Version 1 9-16
 Public transportation was limited in the affected areas.
 Transportation from the National Federation of Public Transportation of Passengers
(Federación Nacional de Cooperativas de Transporte Público de Pasajeros del Ecuador),
FENACOTIP, was used to aid the PN and the firefighters to get to the affected areas.

ONE WEEK AND ONE MONTH AFTER THE EARTHQUAKE

Overall, within a month of the main earthquake the affected areas were serviced by 10,595
and 8,827 dispatched members of the Ecuadorian Armed Forces, and the National Police,
respectively, and 201 firefighters. Their responsibilities were:
 Armed Forces
Security in strategic places and constant security patrols (Fig. 9.6.1).
Administration of the security centers.
Demolition of unsafe structures (Fig. 9.6.2)
Debris removal (Fig. 9.6.2).
Road and transportation maintenance (Fig. 9.6.3).

Figure 9.6.1. An air force soldier stands guards the area in front of collapsed buildings in Manta. Photo
by R. Abd, AP (4/17/16, approximate GPS: 0°57′43″S, 80°42′45″W).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-17
Figure 9.6.2. Demolition and debris removal in Portovejo. Photo by GEER-ATC member R. Gilsanz
(4/28/16, GPS: 1°3'24.61"S, 80°27'17.29"W).

 National Police
Patrols in Portoviejo in the affected area.
Patrols in zone zero.
Patrols in San Vicente.

 Firefighters (Quito Unit)


Coordination with 13 people and 3 vehicles.
Meeting with COE at Cojimies for assessment.

Figure 9.6.3. Army cleanup and restoration operations for road infrastructure. Photo provided to
GEER-ATC team by the Ecuadorian Armed Forces.

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-18
9.7 Fuel and Fires
FUEL
Ecuadorians rely on propane tanks for basic needs like cooking and heating water, with
refills delivered regularly by state-subsidized gas trucks. The earthquake damaged the propane
storage at homes (Fig. 9.7.1) and the propane delivery infrastructure and transportation
systems. Access and purchase of propane was limited or impossible for many families due to
increased prices and lack of availability which affected their livelihood.

Figure 9.7.1. Damaged propane tanks in Pedernales home. Photo by the Ecuadorian Armed Forces.

The Fuel Relief Fund (FRF), is a nonprofit organization comprised of teams of highly
trained and specialized volunteers based in the United States and the Netherlands, with goal to
solve emergency fuel needs in major disasters across the globe. FRF worked with local
officials, the UN and other key humanitarian agencies to identify immediate needs of safe water
and capacity to cook. FRF was on the ground on April 21st and distributed fuel in Manabí using
the most viable transportation routes and vehicles of the local administration (Fig. 9.7.2).

Within a week, with support of donors and participation of local volunteers, 5,000 10-
gallon propane tanks were distributed in Manabí (FRF, 2016). Within the 2nd week, FRF
distributed enough fuel for another 3,625 families, bringing the total number of people reached
to 34,500. In collaboration with the Global Giving donors, an additional 5,045 10-gallon
propane tanks were distributed with fuel for 20,180 people (Fig. 9.7.2). Overall, the distribution
covered the average household’s needs for clean water and cooking for a month (Fig. 9.7.3).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-19
Figure 9.7.2. Fuel Relief Fund propane distribution with local vehicles and volunteers (FRF, 2016).

Figure 9.7.3. Volunteers distribute propane tanks for cleaning the water and cooking (FRF, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-20
FIRES

Since Ecuador does not have a gas or other fuel pipeline network systems, fires after the
earthquake were limited. The few incidents are described below.
On the day of the earthquake, the following fire-related incidents and actions took place:

 Esmeraldas province: The refinery operations ceased due to potential fire alert. One of the
stations reported a fire of an electric transformer that became under control the same day.
 Santo Domingo: The terminal had one fire which was under control.

One day after the earthquake, the following fire-related observations were made:

 Manabí province: A shelter in Pedernales that hosted 9 families was destroyed after an
explosion of a gas tank (Fig. 9.7.4). Firefighters and 10 police officers assisted in
controlling the situation. One person was injured and treated at a nearby ambulance.
 Guayas province: In the southern part of Guayaquil, a three-story house caught fire as a
result of malfunction of the electrical system. No victims or injuries were reported.

Figure 9.7.4. Shelter in Pedernales after gas explosion (El Comercio, 2016).

Two days following the main earthquake event, one house was destroyed by fire in
Portoviejo (Figs. 9.7.5 and 9.7.6) with no victims reported.

A week after the main earthquake event, 7 houses caught on fire in Tosagua, at the province
of Manabí (Fig. 9.7.7). Firefighters from Chone, Calceta, and Junín took control of the
emergency. The same week, investigations started to determine the cause of this fire.

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-21
Figure 9.7.5. House affected by a fire in Portoviejo 2 days after the earthquake (El Universo, 2016).

Figure 9.7.6. Burned belongings in a Portoviejo street 2 days after the event (El Universo, 2016).

Figure 9.7.7. The fire seen in one of the streets of Tosagua (El Telégrafo, 2016; El Diario, 2016).

GEER-ATC Muisne, Ecuador, Earthquake 9-Community Preparedness and Response


Report Version 1 9-22
Blank Page
Blank Page
CHAPTER 10
Conclusions

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
10 Conclusions
The April 2016 Mw7.8 Muisne, Ecuador earthquake and its aftershocks led to hundreds of
fatalities, thousands of injuries, tens of thousands of people homeless, serious damage and
collapses to buildings and infrastructure, and economic impact estimated at 3% of the national
GDP. The performance of the natural and built environment was documented in the
geotechnical, structural and nonstructural aspects by the GEER-ATC reconnaissance team that
visited the most affected areas. The following conclusions have been drawn to date:

1. The ground motions were intense and in some cases exceeded design levels. Factors that
contributed to the intensity of the motions included the difficult subsurface conditions,
which, in the majority of the coastal areas, include site class “F” sites that have either thick
soft and plastic clays or potentially liquefiable loose sands.
2. The extensive structural damage can be attributed to the strong seismic shaking that often
was in resonance with the structural fundamental frequencies, combined with complex
structural irregularities in elevation and plan, and inadequate seismic structural and
nonstructural resistance specifications or construction methods.
3. The infrastructure system suffered extensive structural and nonstructural damage due to the
seismic shaking and liquefaction-induced loss of soil strength and ground deformations.
4. Many healthcare and other critical or essential facilities behaved poorly, often due to
nonstructural damage that interrupted operations critical for the recovery from the
earthquake, with significant social and economic consequences.
5. The generated tsunami was minor and did not induce any significant additional damage to
the coastal areas. However, predictions for future events include substantial tsunami hazard
exposure due to the topographic and tectonic setting.
6. Information distribution to the public could have been more efficient with the use of social
media or television. The people were often confused and did not have a clear understanding
of the measures taken to rebuild their homes or restart their affected businesses.
7. Use of seismic protective technology was beneficial and granted resilient behavior of
transportation facilities, including seismic isolation of the Los Caras Bridge, which
sustained the largest ever recorded bearing displacement, and ground improvement in the
Boca de Briceno Bridge, which mitigated most of the liquefaction effects at the site.

GEER-ATC Muisne, Ecuador, Earthquake 10-Conclusions


Report Version 1 10-1
Overall, the GEER-ATC team effort has yielded a plethora of geotechnical and structural
datasets, lessons, and suggestions for future research. The team is still processing data obtained
during and after the mission, which will be included in the next version of the report. Some
observed short- and long-term needs of Ecuador are:

1. Recently developed local guidelines and construction standards need to be applied in new
construction and retrofit of existing structures. Updates on these guidelines can consider
re-evaluation of the seismic and tsunami hazards based on the 2016 earthquake, and address
the unique site effects that can create large amplification of the ground motion or
liquefaction through microzonation, in-situ testing and studies in the main affected areas.
2. Rebuilding in Ecuador should consider short- and long-term goals and address resiliency
needs for future earthquakes and other hazards that the country is exposed to, such as
tsunamis, landslides, infectious diseases, and climate change. An outreach program can
educate the public and decision makers on the tectonic environment they live in, the hazards
they are exposed to, and the technological tools and actions used by the government and
the engineering community to improve seismic behavior of the built environment.
3. Factors to be considered in future planning may include: (i) better understanding of natural
hazards exposure and associated risks; (ii) knowledge transfer and use of seismic protective
technologies and modern standards; (iii) creation of simple-to-follow, yet technically
accurate protocols for rebuilding essential facilities and testing model houses built with
local construction materials and methods using shaking table experiments to prove their
seismic adequacy and then use at a large scale to host thousands of homeless families; and
(iv) setting resiliency goals to address short-term needs of housing, offering health care and
re-energizing the economy, but also long-term planning in anticipation of future events.

In closing, this devastating earthquake, as many others in the past, reminded us that while
we cannot predict when, where or how big the next earthquake will be, we can anticipate its
potential consequences and use mitigation and protective technology and analytical tools
combined with short- and long-term risk goals to anticipate, respond and recover better when
it happens.

GEER-ATC Muisne, Ecuador, Earthquake 10-Conclusions


Report Version 1 10-2
Blank Page
Blank Page
CHAPTER 11
References

GEER-ATC Muisne, Ecuador 2016


Report Version 1
Blank Page
11 References
Abe, K., 1979. Size of great earthquakes of 1837-1979 inferred from tsunami data, J. Geophys.
Res., 84:1561-1568.
Abrahamson, N., Gregor, N., Addo, K., 2016. BC hydro ground motion prediction equations
for subduction earthquakes, Earthquake Spectra, 32(1):23-44.
Academic, 2016. Ecuador, en.academic.ru/dic.nsf/enwiki/5425, last accessed 9/13/16.
ACAPS, 2016. ACAPS Briefing Note: Earthquake in Ecuador Update,
acaps.org/sites/acaps/files/products/files/160426_acaps_start_briefing_note_ecuador_earthqu
ake_update.pdf, last accessed 9/10/16.
ACI 318-14, 2014. Building code requirements for structural concrete & Commentary,
American Concrete Institute.
ACI 318S-14, 2015. Building code requirements for structural concrete and commentary -
Spanish Language, Requisitos de reglamento para concreto estructural y comentario, versión
en español, American Concrete Institute.
Aguiar, R., 1999. Control del Comportamiento Estructural en Estado Limite del Diseño con
Programa Ceinci2, Centro de Investig. Científicas Escuela Politécnica del Ejército, Ecuador.
Aguiar, R., Varela, M., 1999. El sismo de Bahía, Centro de Investigaciones Científicas Escuela
Politécnica del Ejército, Mendoza, Argentina.
AIS, 2012. Reglamento Colombiano de Construcción Sismo Resistente (NSR-10). Asociación
Colombiana de Ingeniería Sísmica. Bogotá, Colombia.
Aki, K., 1957. Space and time spectra of stationary stochastic waves, with special reference to
microtremors. Bulletin Earthquake Research Institute, 35:415-456.
Amodio, E., 2005. Las furias del temblor. Análisis comparativo de dos sismos históricos:
Quito, febrero 1797 y Cumaná, diciembre 1797, Revista Geográfica Venezolana, 119-141.
Ancheta, T., Darragh, R., Stewart, J., Seyhan, E., Silva, W., Chiou, B., Wooddell, K., Graves,
R., Kottke, A., Boore, D., Kishida, T., Donahue, J., 2013. Pacific Earthquake Engineering
Research, Report PEER 2013/03, PEER NGA-West2 Database.
Andes Newspaper, 2016. 95% of Ecuador’s road system in good condition,
andes.info.ec/en/news/95-ecuadors-road-system-good-condition.html, last accessed 9/13/16.
Andes Newspaper, 2016. Disminuye incidencia del fenómeno El Niño, según estudio del
comité ERFEN de Ecuador, andes.info.ec/es/noticias/disminuye-incidencia-fenomeno-nino-
segun-estudio-comite-erfen-ecuador.html, last accessed 9/7/16.
Andes Newspaper, 2016. Donations and volunteers arrive in Pedernales to help earthquake
victims, Agencia Publica de Noticias del Ecuador y Suramerica, andes.info.ec/donaciones-
voluntarios-llegan-pedernales-auxiliar-victimas-terremoto.html, last accessed 9/23/16.
Arellano, S.R., Hall, M., Samaniego, P., Ruiz, A., Molina, I., Palacios, P., Yepes, H.,
OVTIGEPN, 2008. Degassing patterns of Tungurahua volcano in 1999-2006 period from
remote spectroscopic measurements of SO 2 , J. Volc. & Geoth. Res., 176:151-162.
Argudo J. and Villacrés A., 2001. Manual de diseño sísmico de estructuras para Guayaquil,
Universidad Católica de Santiago de Guayaquil.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-1
Argudo, J., Yela, R., 1995. Vulnerabilidad Estructural de Hospitales de Guayaquil-Ecuador.
ASCE 7-05; 7-10, 2005; 2010. Minimum Design Loads for Buildings and Other Structures,
American Society of Civil Engineers, Codes and Standards Committee
ASCE, 2010. Chapter 13 – Seismic design requirements for nonstructural components, Chapter
13 of ASCE7-10, American Soc. of Civil Engineers - Structural Eng. Institute, Reston, VA.
Asesores y Consultores del Litoral Cia, LTDA (ACOLIT), 2008. Memorias Multipropósito
Chone. Manabí, Ecuador: Provincial Council of Manabí.
Asesoria y Estudios Tecnicos c. ltda, 2010. Complementary Studies for the San Vicente –
Pedernales Route (in Spanish, Estudios Complementarios de la Carretera San Vicente –
Pedernales), Technical Report.
Asesorias y Estudio Tecnico C. LTDA (Undated), Refuerzo de Suelo Usando Pilas de
Agregado Compactado Impact Para Terraplen de Variante Briceno, Informe Technico –
Geopier, Variante Briceno, Technical report.
Ashford, S.A., Sitar, N., 1997. Analysis of topographic amplification of inclined shear waves
in a steep coastal bluff, Bulletin of Seismological Society of America, 87(3):692-700.
Ashford, S.A., Sitar, N., Lysmer, J. and Deng, N., 1997. Topographic effects on the seismic
response of steep slopes. Bulletin of the seismological society of America, 87(3), pp.701-709.
Askew, B., Algermissen, S., 1985. Catalog of S. America earthquakes, hypocenter & intensity
data, Vol. 6 - Ecuador, Centro Regional de Sismologia para America del Sur, Lima, Peru.
ASTM D1586-11, 2011. Standard test method for standard penetration test and split-barrel
sampling of soils, ASTM Int., West Conshohocken, PA.
ASTM D3441-16, 2016. Standard test method for mechanical cone penetration testing of soils,
ASTM Int., W. Conshohocken, PA.
ASTM D6066-11, 2011. Standard practice for determining normalized penetration resistance
of sands for evaluation of liquefaction potential, ASTM Int., W. Conshohocken, PA.
ATC-20, 20-2, 1995. Procedures for post-earthquake safety evaluation of buildings &
Addendum to the ATC-20 post-earthquake building safety evaluation procedures, Applied
Technology Council, Redwood City, CA.
ATC-20-1, 2005. Field manual: post-earthquake safety evaluation of buildings, 2nd edition,
Applied Technology Council, Redwood City, CA.
ATC-20-3, 1997. Case studies in rapid post-earthquake safety evaluation of buildings, Applied
Technology Council, Redwood City, CA.
Athanasopoulos-Zekkos, A. et al., 2016. Liquefaction and geotechnical aspects of the Ecuador
earthquake of 4/16/16, Presentation, mini-symposium: Big picture of infrastructure resilience,
1st Int. Natural Hazards & Infrastructure Conf., Crete, Greece
Avalon Waterways, 2016. Discover the Galápagos Islands, avalonwaterways.com/Experience-
Avalon/Galapagos/, accessed 9/10/16.
Baize, S., Audin, L., Winter, T., Alvarado, A., Moreno, L.P., Taipe M., Reyes, P., Kauffmann,
P., Yepes, H., 2015. Paleoseismology and tectonic geomorphology of the Pallatanga fault
(Central Ecuador), a major structure of the S. American crust, Geomorphology, 237: 14:28

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-2
Baldock, J.W., 1982a. Geology of Ecuador – Bul. of national geological map of Ecuador.
Ministerio de recursos naturales y energéticos, Dirección general de Geología y Minas, 70p.
Baldock, J.W., 1982b. National geological map of Ecuador. Ministerio de Recursos Naturales
y Energéticos, Dirección General de Geología y Minas, Quito, scale 1:1,000,000.
Barberi, F., Coltelli, M., Ferrara, G., Innocenti, F., Navarro, J.M., Santacroce, R., 1988. Plio-
quaternary volcanism in Ecuador. Geological Magazine, 125(1):1-14.
Barrancos, J., Roselló, J.I., Calvo, D., Padrón, E., Melián, G., Hernández, P.A., Pérez, M.,
Millán N., Galle, B., 2008. SO2 emission from active volcanoes measured simultaneously by
COSPEC and mini-DOAS, Pure & Applied Geophysics, 165(1):115-133.
Beauval, C. et al., 2013. An earthquake catalog for seismic hazard assessment in Ecuador,
Bulletin of Seismological Society of America, 103(2A):773-786.
Beauval, C., Yepes, H., Bakun, W. H., Egred, J., Alvarado, A., & Singaucho, J. C., 2010.
Locations and magnitudes of historical earthquakes in the Sierra of Ecuador (1587–1996).
Geophysical Journal International, 181(3):1613-1633.
Begemann, H.K.S., 1963. Use of static penetrometer in Holland, N. Zealand Eng., 18(2):41p.
Bilek, S.L., 2010. Seismicity along South American subduction zone: Review of large
earthquakes, tsunamis, and subduction zone complexity, Tectonophysics, 495(1-2):2–14.
Bolaños, J.E., 2010. Geological and Mining Potential of Ecuador, last accessed 8/1/16.
British Geological Survey, BGS, 2016. Ecuador earthquake disaster response 2016,
bgs.ac.uk/research/earthHazards/epom/ecuadorEarthquake.html, last accessed 9/7/16.
Brookfield, M.E., Hemmings, D.P., Van Straaten, P., 2009. Paleoenvironments and origin of
the sedimentary phosphorites of Napo Formation, S. American Earth Sciences, 28:180-192.
Bryant, J.A., Yogodzinski, G.M., Hall, M.L., Lewicki, J.L., Bailey, D.G., 2006. Geochemical
constraints on the origin of volcanic rocks from the Andean northern volcanic zone, Ecuador.
J. Petrology, 47(6):1147-1175.
Cano Canoscorp, 2016. Terrible fire Tosagua, Manabi, EC (Terrible incendio Tosagua,
Manabí), 4/26/16 online video, youtube.com/watch?v=IFRoXYaqE_k, last accessed 6/11/16.
Capon, J., 1969. High resolution frequency-wavenumber spectrum analysis, IEEE, 57(8):
1408-1418.
Carena, S., 2011. Subducting-plate topography and nucleation of great and giant earthquakes
along the South American trench, Seismological Research Letters, 82(5):629-637.
Carn, S.A., Krueger, A.J. Krotkov, N.A., Arellano, S., Yang, K., 2008. Daily monitoring of
Ecuadorian volcanic degassing from space, J. Volcanology & Geoth. Res., 176(1):141-150.
Castro, A.V.C. 2007. Analysis sismico de estructuras con disipadores de enegria viscoelasticos
usando espectros y accelerogramas, Escuela politecnica del ejercito Carrera de ingenieria civil
scuela, 189 p.
Cavanna, A.J.S., 1965. Earthquake problems of structures in Ecuador, Int. Inst. of Seismology
& Earthquake Engineering, Tokyo, Japan, 2(2):101-128.
Central Intelligence Agency (CIA), 2013; 2015. World Factbook, cia.gov/library/download-
2013/; cia.gov/library/publications/the-world-factbook/, last accessed 9/10/16.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-3
Central Intelligence Agency (CIA), 2016. World Factbook, cia.gov/library/publications/the-
world-factbook/fields/2085.html, last accessed 9/13/16.
Centre for Economic and Policy Research, 2012. Ecuador’s success in emerging from
economic recession; Reducing poverty & unemployment. cepr.net/press-center/ecuadors-
success-emerging-from-economic-recession-reducing-poverty, accessed 9/10/16.
Chávez, M., 2013. Revisión del Diseño de la presa Rio Grande del Proyecto Multipropósito
Chone, Secretaria Nacional del Agua (SENAGUA), Manabí, Ecuador.
Chlieh, M., et al., 2014. Distribution of discrete seismic asperities and aseismic slip along the
Ecuadorian megathrust. Earth & Planetary Science Letters, 400:292-301.
Christopoulos, C., Filiatrault, A., & Bertero, V. V., 2006. Principles of Passive Supplemental
Damping and Seismic Isolation. Iuss press.
Chunga, K., León, C., Quiñónez, M., Benítez, S., Montenegro G., 2002. Seismic hazard
assessment for Guayaquil: Insights from Quaternary Geol. Data, Nat. Geoph. Data Center
Chunga, K., Toulkeridis, T., 2014. First evidence of paleo-tsunami deposits of a major historic
event in Ecuador, J. Tsunami Soc. Int., 33:55-69.
Código Ecuatoriano de la Construcción, 1977. Instituto Ecuatoriano de Normalización, pp 141,
Quito, Ecuador.
Collot, J.-Y. et al., 2004. Are rupture zone limits of great subduction earthquakes controlled by
upper plate structures? Evidence from multichannel seismic reflection data across N Ecuador
– SW Colombia margin, J. Geoph. Res., Solid Earth (1978–2012), 109(B11).
Colmenares, L., Zoback, M.D., 2003. Stress field and seismotectonics of northern South
America. Geology, 31(8):721-724.
Comite del Ano Geofisico Internacional del Ecuador (1959). Breve historia de los principales
teremotos en la republica del Ecuador, Observatorio Astronomico de Quito, Ecuador.
Constantinou, M., Kalpakidisk, I., Filiatrault, A., Ecker, R., 2011. LRFD-based analysis and
design procedures for bridge bearings and seismic isolators, Multidisciplinary Center for
Earthquake Engineering Research, Technical Report MCEER 11-0004, Buffalo, NY.
Constantinou, M., Whittaker, A., Kalpakidis, Y., Fenz, D., Warm, G., 2007. Performance of
seismic isolation hardware under service and seismic loading, Multidisciplinary Center for
Earthquake Engineering Research, Technical Report MCEER 11-0004, Buffalo, NY.
Cox, B., Wood, C., 2011. Surface wave benchmarking exercise: Methodologies, results,
uncertainties, GeoRisk: Geotechnical risk assessment & management (C.H. Juang et al., eds.),
ASCE, GSP 224:845-852.
Crespo, E., O'Rourke, T.D., Nyman, K.J., 1991. Effects on lifelines, Chapter 6, March 5, 1987,
Ecuador earthquakes: mass wasting & socioeconomic effects, National Research Council,
Washington, DC.
Cruz Roja, 2016. Ecuadorian Red Cross distributes 35 mil liters of water a day in Pedernales
(Cruz Roja Ecuatoriana distribuye 35 mil litros de agua diarios en Pedernales), 4/22/16 release,
cruzroja.org.ec/distribuye-35-mil-litros-de-agua-diarios-en-pedernales, last accessed 6/11/16.
Cruz Roja, 2016. Guayaquil Municipality delivers humanitarian help kits to Muisne
(Municipalidad de Guayaquil entrega kits de ayuda humanitaria para atender a comunidades

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-4
de Muisne), 4/23/16 press release by C. Roja, cruzroja.org.ec/boletines-2016/distribuye-35-
mil-litros-de-agua-diarios-en-pedernales, last accessed 6/11/16.
Davila, F., Egtiez, A., 1900. Deformation analysis in the basement of the Chota Basin:
Ambaqui Group, Bol. Geol. Ecuatoriano, 1(1):39-52.
De Ruiter, J., 1971. Electric penetrometer for site investigations, ASCE J. Soil Mechanics
& Foundations, 97(SM2):457-472.
De Ruiter, J., 1981. Current penetrometer practice, in cone penetration testing & experience,
ASCE Geot. Eng. Div., St. Louis, MI, 1-48.
Dialogo Americas, 2016. Ecuadorian armed forces deploy 10,000 soldiers to help quake
victims, Digital Military Magazine, 4/19/16 article by H. Alava, dialogo-
americas.com/articles/ecuadorian-armed-forces-deploy-10000-soldiers-help-quake-victims,
last accessed 6/12/16.
Diaz-Fanas, G., Nikolaou, S., Vera-Grunauer, X., Gilsanz, R., Diaz, V., 2016. Observations
from the 4/16/16 Muisne, Ecuador earthquake, Keynote paper, mini-symposium: Big picture
of infrastructure resilience, 1st Int. Natural Hazards & Infrastructure Conf., Crete, Greece
DJI, 2016a. PHANTOM 4, dji.com/phantom-4, last accessed 9/8/16.
DJI, 2016b. INSPIRE 1 PRO/RAW, dji.com/inspire-1-pro-and-raw , last accessed 9/8/16.
Dumont, J.F., Santana, E., Vilema, W., 2005. Morphologic evidence of active motion of the
Zambapala fault, Gulf of Guayaquil, Geomorphology, 65(3):223-239.
Dziewonski, A.M., Chou, T.-A, Woodhouse, J.H., 1981. Determination of earthquake source
parameters from waveform data of global/regional seismicity, J. Geoph. Res., 86:2825-52.
ECHO, 2016. Floods/rainy season (ECHO, Ecuador Risk Management Secretariat, Redhum,
media) (ECHO Daily Flash, 4/14/16), reliefweb.int/report/ecuador/, last accessed 9/7/16.
Ecuador Army Corps of Engineers, 2010. Structural and design drawings for Bahia Bridge.
Ecuador Army Corps of Engineers, 2012. Information from www.cee.gov.ec/.
Ecuador Ministerio de Transporte y Obras Publicas (EMTOP), 2010a. Design drawings
containing Complimentary studies for the San Vicente- Pedernales road (in Spanish Estudios
Complementarios para la Carretera San Vicente – Pedernales Implantacion General), June.
Ecuador Ministerio de Transporte y Obras Publicas (EMTOP), 2010b, Implantacion General,
Puente sobre el Rio Briceno, June.
Ecuador Ministerio de Transporte y Obras Publicas (EMTOP), 2011. Design drawing contract
No. 008-2011, August.
Ecuador Ministerio de Transporte y Obras Publicas (EMTOP), 2011. Planos Estructurales,
Complementacion de estudios para la ampliacion, rectificacion y mejoramiento de la carretera
Portoviejo-Crucita en la Provincia de Manabi, August.
Ecuador Ministerio de Transporte y Obras Publicas (EMTOP), 2014. Implantacion General,
Puente sobre el Rio Briceno.
Egbue, O., Kellogg, J., 2010. Pleistocene to present North Andean “escape,” Tectonophysics,
489(1):248-257
Egred J., 2000. El terremoto de Riobamba, Abya-Yala, Riobamba, 2:107 pp.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-5
Egred J., 2009. Catalogo de terremotos del Ecuador 1541-2009, Escuela Politecnica Nacional,
Instituto Geofisico, Internal report, in spanish.
Ekström, G., Nettles, M., Dziewonski, A.M., 2012. The global CMT project 2004-10:
Centroid-moment tensors for 13,017 earthquakes, Phys. Earth Planet. Int., 200-201:1-9.
El Comercio Newspaper, 2016. Cards or vouchers to improve food delivery to earthquake
victims (Con tarjetas o vouchers se busca mejorar entrega de alimento a damnificados por el
terremoto), 4/27/16 online article by M. Orozco, elcomercio.com/actualidad/tarjetas-vouchers-
alimentos-terremoto-ecuador.html, last accessed 6/13/16.
El Comercio Newspaper, 2016. Fire destroys a house in Pedernales inhabited by earthquake
victims (Incendio destruye una casa en Pedernales donde vivían damnificados), 5/17/16 article
by A. Veintimilla, elcomercio.com/actualidad/incendio-pedernales-damnificados-terremoto-
ecuador.html, last accessed 6/9/16.
El Comercio Newspaper, 2016. Fire in 7 houses scares again the people of Tosagua (Un
incendio de 7 casas volvió a asustar a la población de Tosagua), 4/26/16 article by M. Gonzales,
elcomercio.com/actualidad/incendio-casas-tosagua-emergencia-ecuador, last accessed 6/9/16.
El Comercio Newspaper, 2016. Help is not limited to donations (La ayuda tras el terremoto no
se limita a las donaciones), 4/26/16 article by A. Medina, elcomercio.com/ayuda-donaciones-
terremotoecuador-empresas-economicas, last accessed 6/9/16.
El Comercio Newspaper, 2016. Three more cases of Zika in Ecuador, rising to 56. (in Spanish,
Tres casos mas de zika en Ecuador, suben a 56 los registros), elcomercio.com/actualidad/casos-
zika-ecuador-aedes-enfermedad.html, online 2/25/16 article, accessed 9/10/16.
El Diario Newspaper, 2016. Voracious fire consumes 6 houses in Tosagua (Voraz incendio
consume 6 viviendas en Tosagua), 4/26/16 online article, eldiario.ec/noticias-manabi-
ecuador/389843-voraz-incendio-consume-seis-viviendas-en-tosagua/, last accessed 6/9/16.
El Telégrafo Newspaper, 2016. Six houses destroyed by fire in Tosagua (6 casas fueron
destruidas por un incendio en Tosagua), 4/26/16 article, eltelegrafo.com.ec/noticias/regional-
manabi/1/seis-casas-fueron-destruidas-por-un-incendio-en-tosagua, last accessed 6/9/16.
El Universo Newspaper, 2016. Manta, Portoviejo, Manabi towns with extensive damage after
earthquake (Manta, Portoviejo y más localidades manabitas con graves daños tras terremoto),
4/16/16 article by J.G. Espinel, eluniverso.com/noticias/2016/04/16/nota/5527542/manta-
portoviejo-mas-localidades-manabitas-graves-danos , last accessed 6/10/16.
El Universo, 2012. 80% of Ecuatorians is Catholic, according to INEC (in Spanish, El 80% de
los ecuatorianos afirma ser católico, según el INEC), eluniverso.com/2012/08/15/1/1382/80-
ecuatorianos-afirma-ser-catolico-segun-inec.html, accessed 9/10/16.
ESPE-Innovativa & Empresa Publica del Agua, 2015. Instrumentación, Proyecto Propósito
Múltiple Chone, Fase 1. Manabí, Ecuador.
Feininger, T., Bristow, C.R., 1980. Cretaceous and Palaeocene geological history of coastal
Ecuador, Geol, Res., 69:849-874.
FEMA E-74, 2005. Reducing the Risks of Nonstructural Earthquake Damage: A Practical
Guide, Federal Emergency Management Agency, Washington, D.C.
FEMA P-695, 2009. Quantification of building seismic performance factors, Federal
Emergency Management Agency, Washington, D.C.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-6
Finn, W.L., Ruz, F., 2015. Amplification effects of thin soft surface layers: A study for NBCC-
15. Perspectives on Earthquake Geotechnical Eng. in Honour of Prof. Ishihara, A. Ansal &
M. Sakr, eds., Springer, 37:33-43.
Flyg London, 2011. Los Caras Bridge: A giant step for development for Ecuador,
flyglondon.com/los-caras-bridge, last accessed 9/19/16.
Font, Y., Segovia, M., Vaca, S., Theunissen, T., 2013. Seismicity patterns along the Ecuadorian
subduction zone: new constraints from earthquake location in a 3D a priori velocity model,
Geophysical J. Int., 193(1):263-286.
Food & Agriculture Organization (FAO) of United Nations, 2005. National Aquaculture
Sector Overview Fact Sheets, Ecuador. by Schwarz, L. in: FAO Fisheries & Aquaculture
Department, Rome, updated February 1.
Foti, S., Lai, C., Rix, G., Strobbia, C., 2014. Surface wave methods for near-surface site
characterization, CRC Press, Boca Raton, FL.
Freymuller, J.T., Kellogg, J.N., Vega, V., 1993: Plate motions in the north Andean region. J.
Geophys. Res., 98:21853-21863.
Fuel Relief Fund (FRF), 2016. Ecuador Earthquake, globalgiving.org/Ecuador_Report.pdf, last
accessed 9/23/16.
Gallagher, J.J., 1989. Andean chronotectonics, Geology of the Andes and its relation to
hydrocarbon and mineral resources, Earth Sci. Ser., G.E. Erickson et al., eds., 11:23-37.
García, L.E., 1984. Development of the Colombian Seismic Code, 8th World Conference on
Earthquake Engineering, Earthquake Engineering Research Institute, EERI, 747-754.
Garcia, L.E., Perez A., Bonacci J., 1996. Cost implications of drift controlled design of
reinforced concrete buildings, 11th World Conf. Earthq. Eng., Elsevier, Paper No. 548.
Garcia-Aristizabal, A. et al., 2007. Seismic, petrologic, geodetic analyses of the 1999 dome-
forming eruption of Guagua Pichincha volcano, Ecuador. J. Volc. Geoth. Res., 161:333-351.
GEER, 2016. Manual for GEER Reconnaissance Teams, v.4, Geotechnical Extreme Events
Reconnaissance Association, geerassociation.org/Important%/GEER_ Manual_2014_v4.pdf,
last accessed 4/20/16.
©2013-2015.
GEO Salud Ministerio de Salud Publica, aplicaciones.msp.gob.ec/
/dniscg/geosalud/gui/#, last accessed 9/20/2016.
Geoestudios, SA, 2014. Seismic and geotechnical microzonation of Guayaquil in accordance
to EC Construction Norm (NEC-15), Report to Risk Management Secretary, Ecuador.
Geoestudios, SA, 2016. Measurement of shear wave velocity in strong motion stations from
Ecuador, Data collected for GEER, Geoestudios, SA.
Geoestudios, SA, 2016. Post-earthquake borings compilation.
Giesecke, A., Capera A.A.G., Leschiutta I., Migliorini E., Valverde L.R., 2004. CERESIS
earthquake database of Andean Region, Annals of Geophysics, 47(2/3):421-435.
Gilsanz, R., Vera-Grunauer, X., Nikolaou, S. Miranda, E., Diaz, V., Hernandez, L., Arteta
Torrents, C., Diaz-Fanas, G., Luque, R, to be published in 2017. Structural Observations - M7.8
- April 16, 2016 Ecuador Earthquake; 16th World Conference on Earthquake Engineering,
16WCEE 2017.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-7
Google Inc., 2016. Google Earth Pro (Version 7.0.2.8415) [Software]. Mountain View, CA.
Green Coconut Rum, 2016. Galapagos Islands, greencoconutrun.com/our-route/season-
2/galapagos-islands/, accessed 9/10/16.
Gusiakov, V.K., 2005: Tsunami generation potential of different tsunamigenic regions in the
Pacific. Marine Geology, 215(1-2):3-9.
Gutscher, M.-A., Malavieille, J., Collot, J.-Y., 1999. Tectonic segmentation of North Andean
margin: impact of Carnegie ridge collision. Earth Planet. Sci. Lett., 168(3):255–270.
Hall, M. L., Samaniego, P., LePennec, J.L., Johnson, J.B., 2008. Ecuadorian Andes volcanism:
A review of Late Pleistocene to present activity. J. Volcanol. & Geoth. Res., 176(1):1-6.
Harpp, K.S., Fornari, D.J., Geist, D.J., Kurz, M.D., 2003. Genovesa Submarine Ridge: A
manifestation of plume‐ridge interaction in the northern Galápagos Islands, Geochemistry,
Geophysics, Geosystems, 4(9).
Haymon, R.M., et al., 1993. Volcanic eruption of the midocean ridge along East Pacific Rise
crest at 9°45–52′N: Direct submersible observations of seafloor phenomena associated with an
eruption event in April, 1991, Earth Planet. Sci. Lett., 119(1–2):85–101.
Herd, D.G., Youd, T.L., Meyer, H., Arango C., Waverly J.P., Mendoza, C., 1981. The great
Tumaco, Colombia earthquake of 12 December 1979, Science, 211(4481):441-445.
Hey, R., 1977. Tectonic evolution of the Cocos-Nazca spreading center, Geological Society of
America Bulletin, 88(10):1404–1420.
Holden, J.C., Dietz, R.S., 1972. Galápagos gore, NazCoPac triple junction and Carnegie/Cocos
ridges. Nature, 235:266-269.
Holloway, H.L, 1932. Gold of Ecuador, Mining Magazine, 46:219-223.
IBC 09; 13, 2009, 2013. International Building Code, Int. Code Council ICC, F. Church, VA.
Instituto Geofísico at the Escuela Politécnica Nacional, IG-EPN, 2016a. Acelerógrafos,
Publicado en Instrumentación, igepn.edu.ec/acelerografos, last accessed 7/1/16.
Instituto Geofísico at the Escuela Politécnica Nacional, IG-EPN, 2016b. Informes de los
ultimos sismos, igepn.edu.ec/portal/ultimo-sismo/informe-ultimo-sismo, last accessed 7/8/16.
Instituto Oceanográfico de la Armada, INOCAR, 2016. Informe técnico de tsunami 16 -abril-
2016, online report, inocar.mil.ec/web/index.php/institucion/resena-historica/36-informes-
tecnicos/589-informe-tecnico-de-tsunami-16-abril-2016, last accessed 7/1/16.
International Food Policy Research Institute, 2015. 2015 Nutrition country profile: Ecuador,
ifpri.org/publication/2015-nutrition-country-profile-ecuador, accessed 9/10/16.
International Seismological Centre, ISC, 2015. Global Instrumental Earthquake Catalogue
ISC-GEM v2 (1900-2009).
Ioualalen, M., Ratzov, G., Collot, J.Y., Sanclemente, E., 2011. The tsunami signature on a
submerged promontory: Atacames Promontory, Ecuador. Geoph. J. Int., 184(2):680-688.
Jaillard, E. et al., 1995. The role of Tethys in the evolution of N. Andes between Late Permian
and Eocene times, The Ocean Basins and Margins, Nairn et al., eds., 8:463-491.
Jaillard, E., Soler, P., G. Carlier, G., Mourier, T. 1990. Geodynamic evolution of northern and
central Andes during early to middle Mesozoic: A Tethyan model, Geol. Soc., 147:1009-1022.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-8
Joh, S.H., 1996. Advances in interpretation and analysis techniques for spectral-analysis-of-
surface-waves (SASW) measurements. Ph.D. Dissertation, University of TX, Austin.
Johnson, G.L., Lowrie, A., 1972: Cocos and Carnegie Ridges result of the Galapagos “hot
spot”?, Earth & Planetary Science Letters, 14(2):279-280.
Kanamori, H., 1977. The energy release in great earthquakes, J. Geoph. Res., Solid Earth
(1978–2012), 82(20):2981–2987.
Kanamori, H., McNally, K.C., 1982. Variable rupture mode of the subduction zone along the
Ecuador-Colombia coast, Bull. Seismological Society of America, 72(4):1241-1253.
Kelleher, J.A., 1972. Ruptures zones of large South American earthquakes and some
predictions. J. Geoph. Res., 77(11):2087-2103.
Kellogg, J.N., Vega, V., 1995. Tectonic development of Panama, Costa Rica and Colombian
Andes: Constraints from GPS studies and gravity. Geol. Soc. Am. Sp. Paper 295:75–90.
Kennerley, J.B., 1980. Outline of Geology of Ecuador. Overseas Geology Resources, 55.
Kunstaetter, R., Kunstaetter, D., 2007. Footprint Ecuador and Galápagos, Footprint Travel
Guides, 6th ed., 244pp.
La Info, 2014. Parliament passed Ecuadorian social security agreement with Colombia,
lainfo.es/2014/parliament-passed-ecuadorian-social-security-agreement-with-colombia/,
published 5/21/14, accessed 9/10/16.
Laboratorio de Mecánica de Suelos, Hormigones y Asfaltos LUP, 2013. Soil Study for Bridge
over Rio Chico in Prtoviejo, Manabi, Ecuador (in Spanish, Estudio de Suelo del Puente Sobre
el Río Chico del Canton Portoviejo Provincia de Manabi - Ecuador), Technical Report.
Lebrat, M., Megard, F., Dupuy, C., 1986. Preorogenic volcanic assemblages and the position
of the suture between oceanic terrains and the S. American continent in Ecuador, Zentralblatt
der Geologische Palaontologische Tesl., 1(9/10):1207-1214.
Lees, J.M. et al., 2008. Reventador volcano 2005: Eruptive activity inferred from seismo-
acoustic observation. J. Volcan. & Geoth. Res., 176(1):179-190.
Life Magazine, 1949. Disaster strikes Ecuador: In the shadows of the Andes 5,000 die in
earthquake, August 22.
Lifewithoutlimbs, 2013. Policia and Presidentes in Quito, lifewithoutlimbs.org/nicks-
blog/policia-and-presidentes-in-quito/, accessed 9/10/16.
Litherland, M. et al., 1990. The geology and mineral potential of the Cordillera Real, Ecuador.
Summary of 1986-1990 Geological Project, INEMIN/Mision Britannica, BGS Open File Rep.
Litherland, M., Aspden, J.A., 1992. Terrane boundary reactivation: A control on the evolution
of the northern Andes. South American Earth Sciences, 5(1):71–76.
Litherland, M., Aspden, J.A., Jemielita, R.A., 1994. The metamorphic belts of Ecuador,
Overseas Memoir 11, British Geological Survey, 147 p.
Litherland, M., Zamora, A., 1991. A terrane configuration for the northern Andes, Commun.
Univ. Chile, Santiago, 42:122-126.
Litherland, M., Zamora, A., Egtiez, A. 1993. Mapa geologico de la Republica del Ecuador,
scale 1:1,000,000, British Geologic Surveying, Quito.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-9
Lonsdale, P., 2005. Creation of Cocos and Nazca plates by fission of the Farallon plate,
Tectonophysics, 404(3-4):237–264.
Louie, J.N., 2001. Faster, better shear wave velocity to 100 m depth from refraction
microtremor arrays, Bulletin of Seismological Society of America, 91(2):347-364.
Mac Leod, M. J., Knapp, G. W., Velez, H. P., 2016. Ecuador, Encyclopædia Britannica,
britannica.com/place/Ecuador, accessed 9/10/16.
Maraillou, B., Spence G., Collot J.-Y., Wang K., 2006. Thermal regime from bottom
simulating reflectors along N. Ecuador - S. Colombia margin: Relation to margin segmentation
& great subduction earthquakes, J. Geoph. Res., 111(B12407).
Mayne, P.W., Christopher, B.R., DeJong, J., 2002. Subsurface investigations, Geotechnical
site characterization, Federal Highway Admin., FHWA NHI-01-031, US Dept. of
Transportation Course No. 132031, Washington DC, 300 p.
McCormick, B.T. et al., 2014. A comparison of satellite‐ and ground‐measurements of SO2
emissions from Tungurahua volcano, J. Geoph. Res., Atmosph., 119(7):4264-4285.
McGregor, J.A., Duncan, J.M., 1998. Performance and use of the SPT in geotechnical
engineering practice, Virginia Polytechnic Institute, Blacksburg, VA.
MéGard, F., 1987. Cordilleran and marginal Andes: Review of Andean geology north of Arica
Elbow, S. Circum-Pacific Orogenic Belts & Evolution of Pacific Ocean Basin, 71-95.
Mendoza, C., Dewey, J.W., 1984. Seismicity associated with the great Colombia-Ecuador
earthquakes of 1942, 1958, 1979: Implications for barrier models of earthquake rupture,
Bulletin of Seismological Society of America, 74(2):577–593.
Metro Ecuador Newspaper, 2016. Fire under control south of Guayaquil (Incendio controlado
al sur de Guayaquil), 4/17/16 online article, metroecuador.com.ec/noticias/incendio-
controlado-al-sur-de-guayaquil/rUrpdr---LlTyzbwZ2dIM/, last accessed 6/9/16.
Michaud, F., Witt, C., Royer, J.Y., 2009. Influence of subduction of Carnegie volcanic ridge
on Ecuadorian geology: Reality & fiction, Geol. Soc. America Memoirs, 204(0):217–228.
Miranda, E., Acosta, A., Ceferino, L., Luque, R., 2016. Two-month anniversary of April 16
Mw7.8 Ecuador earthquake: Post-earthquake reconnaissance, PEER Seminar, Univ. CA,
Berkeley, youtube.com/, last accessed 6/19/16.
Monzier, M., Robin, C., Samaniego, P., Hall, M.L., Cotten, J., Mothes, P., Arnaud, N., 1999.
Sangay volcano, EC: structural development, present activity, petrology. J. Volcanology &
Geoth. Res., 90(1):49-79.
Müller, R.D., Roest, W.R., Royer, J.-Y., Gahagan, L.M., Sclater, J.G., 1997. A digital age map
of the ocean floor, J. Geophysical Research, 102(3):211-214.
Municipal GAD Guayaquil, 2014. guayaquil.gob.ec/, accessed 9/10/16.
National Geophysical Data Center, 2012. Comments for the significant earthquake; Comments
for the tsunami event, online.
National Institute for Statistics and Census of Ecuador, 2010. Censo 2010. Instituto Nacional
de Estadística y Censos. 2010. Retrieved 12/13/11.
National Oceanic and Atmospheric Administration, NOAA, 2016. Natural Hazards,
ngdc.noaa.gov/, last accessed 6/15/16.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-10
Nature Communications, 2015. Unravelling the hidden ancestry of American admixed
populations, nature.com/articles/ncomms7596#supplementary-information, accessed 9/10/16.
NBC News (2016). Ecuador earthquake: State of emergency declared after at least 272 killed,
4/17/16 online article by P. Helsel, M. Grimson, E. Fieldstadt of Associated Press & Reuters,
nbcnews.com/news/world/ecuador-earthquake-state-emergency-declared-after-least-272-
killed-n557181, last accessed 5/29/16.
NEC-15, 2015. Norma Ecuatoriana de la Construcción, Ministerio de Desarrollo Urbano y
Vivienda, MIDUVI, Quito.
NEHRP, 2015. Recommended seismic provisions for new buildings and other structures,
National Earthquake Hazards Reduction Program, Federal Emergency Agency, FEMA.
New York Times, 1987. 300 reported dead and 4,000 missing in Ecuador quakes, Special
report to the New York Times, L. Chavez, March 12.
Newhall, C.G., Self, S., 1982. The Volcanic Explosivity Index (VEI): An estimate of explosive
magnitude for historical volcanism, J. Geophysical Research, 87(C2):1231-1238
Nikolaou et al., 2012. Geo-Seismic Design in Eastern US: State of Practice, ASCE
GeoCongress, Keynote State of Practice paper and Presentation, March 25-29, Oakland, CA.
Nikolaou, S., Gilsanz, R., et al., 2015. Learning from structural success rather than failures,
Structure Magazine, Performance issues on extreme events, March:25-13.
NOAA Center for Tsunami Research, NCTR 2016. Ecuador Tsunami, April 16, 2016, online
research report, nctr.pmel.noaa.gov/ecuador20160416/, last accessed 7/1/16.
Nocquet, J.M. et al., 2014. Motion of continental slivers and creeping subduction in the
northern Andes, Nature Geosc., 7:287–291.
Nocquet, J.M. et al., 2016. Shallow earthquake inhibits unrest near Chiles-Cerro Negro
volcanoes, Ecuador-Colombian border, Earth & Planetary Science Letters, 450.
Nogoshi, M., Igarashi, T., 1971. On the amplitude characteristics of microtremor, Part 2 (in
Japanese with English abstract), Seism. Soc. Japan, 24:26-40.
Nuques y Luque Ingenieros Consultores (NYLIC), 2001. Borings at Mejia Bridge, Portoviejo-
Tosagua Road (in Spanish, Sondeos Via Portoviejo – Tosagua, Puente Mejia), October.
OCHA, 2016. REDLAC: Latin America and Caribbean weekly note on emergencies, 4/4/16,
reliefweb.int/sites/reliefweb.int/files/resources/LAC-Report-Weekly_Note_On_
Emergencies-ROLAC-201600404_ENG-20160404-MR-17998.pdf, last accessed 9/7/16.
Organization of American States, 2011. Legal system of Ecuador, oas.org/juridico/
/ecu/en_ecu-int-description.pdf, accessed 9/10/16.
Oromar TV, 2016. Medical services at hospital Rodriguez Zambrano, Manta (Atención Médica
en el hospital R. Zambrano de Manta), 4/24/16 online video, youtube.com/watch?v=Z9-
NlYZ6BIY, last accessed 6/9/16.
Pararas-Carayannis, G., 1980. The earthquake and tsunami of December 12, 1979, Colombia,
Int. Tsunami Inf. Center Report, Abstracted article in Tsunami Newsletter, XIII(1).
Pararas-Carayannis, G., 2012. Potential of tsunami generation along Colombia/Ecuador
subduction & Dolores-Guayaquil mega-thrust, Science, Tsunami Hazards, 31(3):209-230.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-11
Pardo-Casas, F., Molnar, P., 1987. Relative motion of the Nazca (Farallon) and South America
plate since Late Cretaceous time, Tectonics, 6:233-248.
Park, C.B., Miller, R.D., 2008. Roadside passive multichannel analysis of surface waves
(MASW), J. Environmental and Engineering Geophysics, 13(1):1-11.
Park, C.B., Miller, R.D., Ryden, N., Xia, J., Ivanov, J., 2005. Combined use of active and
passive surface waves. J. Engineering & Environmental Geophysics, 10(3):323-334.
Park, C.B., Miller, R.D., Xia, J., 1999. Multichannel analysis of surface waves, Geophysics,
64(3):800-808.
Pilger, R.H., 1983. Kinematics of South American subduction zone from global plate
reconstructions. Geodynamics of eastern Pacific, Caribbean & Scotia arcs: 113-125.
Pontoise, B. and Monfret, T. (2004). Shallow seismogenic zone detected from an off-onshore
temporary seismic network in Esmeraldas, Geochemistry, Geophysics, Geosystems, 5(2).
Ramos, V. A., 2009. Anatomy and global context of the Andes: Main geologic features and
the Andean orogenic cycle. Geological Society of America Memoirs, 204: 31-65.
Ratzov, G., Collot, J. Y., Sosson, M., Migeon, S., 2010. Mass-transport deposits in northern
Ecuador subduction trench: Result of frontal erosion over multiple seismic cycles. Earth &
Planetary Science Letters, 296(1):89-102.
Ratzov, G., Sosson, M., Collot, J. Y., Migeon, S., Michaud, F., Lopez, E., LeGonidec, Y.,
2007. Submarine landslides along N. Ecuador - S. Colombia convergent margin: possible
tectonic control, Submarine Mass Movements & Their Consequences, Springer, 47-55
Ray, R. & Kozameh, S. 2012. Ecuador's Economy Since 2007, May 2012, p. 15.
Rehabilitation Center, Province of Manabí, Republic of Ecuador, 1983. Ajustes al Diseño de
la Presa La Esperanza del Proyecto Carrizal-Chone. Manabí, Ecuador.
Restrepo, J.D., Kjerfve B., 2000. Water discharge and sediment load from western slopes of
Colombian Andes with focus on Rio San Juan, J. Geology, Univ. of Chicago, 108:17-33.
Reynaud, J.Y., Tessier, B., Proust, J.N., Dalrymple, R., Bourillet, J.F., De Batist, M.,
Lericolais, G., Berne, S. Marsset, T., 1999. Architecture and sequence stratigraphy of a late
Neogene incised valley at shelf margin, southern Celtic Sea, J.Sedimentary Research, 69(2).
Ridolfi, F., Puerini, M., Renzulli, A., Menna, M., Toulkeridis, T., 2008. The magmatic feeding
system of El Reventador volcano constrained by texture, mineralogy and thermobarometry of
the 2002 erupted products, J. Volcanology & Geothermal Research, 176(1):94-106.
Ripalda, O., 2007. Boring logs for Port of Manta Expansion.
Ripalda, O., 2016. Analysis of landslides in the San Isidro area in Manabi (in Spanish, Analisis
De Deslizamientos De Tierra Ocurridos En Sector San Isidro, Provincia De Manabi).
Robin, C., Samaniego, P., Le Pennec, J. L., Mothes, P., Van Der Plicht, J., 2008. Late Holocene
phases of dome growth and Plinian activity at Guagua Pichincha volcano, J. Volcanology &
Geothermal Research, 176(1):7-15.
Romo, M., 2016. Notes and Data collected post-earthquake April 2016, Ecuador’s Army Corps
of Engineers team.
Rudolph E., Szirtes S., 1911. Das kolumbianische Erdbeben am 31 Januar 1906, Gerlands
Beitr. z. Geophysik, 2:132- 275.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-12
Samtani, N.C. and Nowatzki, E.A., 2006. Soils and Foundations: Reference Manual, Vol. 1,
Publication No. FHWA-NHI-06-088, National Highway Institute, Federal Highway
Administration, Washington, D.C.
Schmertmann, J.H., Palacios, A., 1979. Energy dynamics of SPT, ASCE, J. Geotechnical
Engineering Division, 105(8):909-926.
Schuster, R., Bolton, P., Comfort, L., Crespo, E., Nieto, A., Nyman, K., O’Rourke, T., 1991.
The March 5, 1987, Ecuador earthquakes: Mass wasting and socioeconomic effects, National
Academies Press, 184 pp.
Schuster, R., Highland, L.M., 2001. Socioeconomic & envir. impacts of landslides in w.
hemisphere, USGS 01-0276, pubs.usgs.gov/of/2001/ofr-01-0276/, last accessed 5/3/16.
Secretaría de Gestión de Riesgos, 2016. Sit. Rep. No. 4 (16/4/16) 24h00 Mw7.8 Muisne
Earthquake (Informe de situación No. 4 (16/4/16) 24h00 Terremoto 7.8° Muisne), 4/16/16
article, Risk Secreatariat, gestionderiesgos.gob.ec/wp-content/downloads/2016/04/Informe-
de-Situaci%C3%B3n-4-24h00.pdf, last accessed 7/15/16.
Secretaría de Gestión de Riesgos, 2016. Situation Rep. No. 66 (18/05/2016) 17h30 Mw7.8
Pedernales Earthquake (Informe de situación No. 66 (18/05/2016) 17h30 Terremoto 7.8°
Pedernales), 4/16/16 article by Risk Secreatariat, gestionderiesgos.gob.ec/wp-content/
downloads/2016/05/Informe-No.-66-9plica-68-18052016-11h00.pdf , last accessed 7/15/16.
Secretaría Nacional de Planificacion y Desarrollo (SENPLADES), 2016. Sp. Ed. Official
Registry No. 290 (Registro Oficial Edición Especial N°290), 5/28/12 online article by
Secretary F. Benítez, planificacion.gob.ec/wp-content/downloads/2013/05/REGISTRO-
OFICIAL_DISTRITOS-Y-CIRCUITOS, last accessed 7/13/16.
Segovia, M., Pacheco, J., Shapiro, N., Yepes, H., Guillier, B., Ruiz, M., Calahorrano, A.,
Egred, D., 1999. The August 4, 1998, Bahia earthquake Mw7.1: rupture mechanism and
comments on the potential seismic activity, 4th ISA G. Goettingen (Gernuinv), 673-677.
Servicio Integrado De Seguridad 911, 2016. Videovigilancia del ECU 911 permite monitorear
zona de desastre tras sismo, dated 4/20/16, ecu911.gob.ec/videovigilancia-del-ecu-911-
permite-monitorear-zona-de-desastre-tras-sismo/, last accessed 7/18/16.
SESAME, 2004. Guidelines for implementation of H/V spectral ratio on ambient vibrations:
measurements, processing, interpretation, sesamefp5.ujfgrenoble.fr/Del-D23, 62 p.
Shepperd, G.L., Moberly, R., 1981. Coastal structure of the continental margin, northwest Peru
and southwest Ecuador. Geological Society of America Memoirs, 154: 351-392.
Singaucho J.C., Laurendeau A., Viracucha C., Ruiz M., 2016. Observaciones del sismo del 16
de Abril de 2016 de magnitud Mw7.8, Intensidades y Aceleraciones, Sometido a la Revista
Politécnica.
Spikings, R.A., Seward, D., Winkler, W., Ruiz, G.M., 2000. Low-temperature thermos-
chronology of the northern Cordillera Real, Ecuador, Tectonics, 19(4):649-668.
Steffke, A.M., Fee, D., Garces, M., Harris, A., 2010. Eruption chronologies, plume heights and
eruption styles at Tungurahua volcano: Integrating remote sensing techniques and infrasound.
J. Volcanology & Geothermal Research, 193(3):143-160.
Storchak, D.A., DiGiacomo, D., Bondár, I., Engdahl, E.R., Harris, J., Lee, W.H.K., Villaseñor,
A., Bormann, P., 2013. Public Release of ISC-GEM Global Instrumental Earthquake Catalogue
(1900-2009), Seism. Res. Lett., 84(5):810-815.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-13
Strasser, F.O., Arango, M.C., Bommer, J.J., 2010. Scaling of the source dimensions of interface
and intraslab subduction-zone earthquakes with moment magnitude, Seism. Res. Letters,
81(6):941-950.
Subterra Ingenieria Geotecnica, 2016. Post-earthquake CPT Compilation.
Swenson, J.L., Beck, S.L., 1996. Historical 1942 Ecuador & Peru subduction earthquakes, and
cycles along Colombia–Ecuador & Peru segments. Pure Appl. Geophys., 146(1):67–101.
Telesur, 2016. Ecuador: Fire registered as a result of earthquake (Spanish: se registra incendio
en Guayaquil como consecuencia del sismo), 4/18/16 online video, msn.com/es-
xl/video/noticias/ecuador-se-registra-incendio-en-guayaquil-como-consecuencia-del-
sismo/vp-BBrV6Lm , last accessed 6/9/16.
Terra Hidro servicios y construcciones S.A. (TH), 2013. Boring logs extracted from Revision
of Rio Grande Dam Design for Multipurpose Chone – Senagua Project (in Spanish, Revision
del Diseño de la Presa Río Grande del Proyecto Multiproposito Chone – Senagua).
The Watchers, 2016. Torrential rainfalls trigger flooding and landslides in Ecuador, article by
E. Ulgrin dated 2/4/16, watchers.news/2016/02/04/torrential-rainfalls-trigger-flooding-and-
landslides-in-ecuador/, last accessed 9/7/16.
The World Economic Forum’s Centre for Global Competitiveness and Performance, 2015. The
Global Competitiveness Report 2015, knoema.com/WFGCI2015/the-global-competitiveness-
report-2015, last accessed 9/13/16.
Tokimatsu, K., 1997. Geotechnical site characterization using surface waves, 1st Int. Conf.
Earthquake Geotechnical Engineering, Ishihara, ed., Balkema, 3:1333-368.
Tokimatsu, K., Shinzawa, K., Kuwayama, S., 1992. Use of short-period microtremors for Vs
profiling, ASCE J. Geotechnical Engineering, 118(10):1544-1558.
Toulkeridis, T. et al. (2016) The 7.8 Mw Earthquake and Tsunami of the 16th April 2016 in
Ecuador - A seismic evaluation and geological field survey, under review for publication.
Toulkeridis, T., 2007. The summer 2006 volcanic crisis of Tungurahua, Ecuador: No Lessons
Learned, AGU Spring Meeting Abstracts, 1:02.
Toulkeridis, T., 2011. Volcanic Galápagos Volcánico. Ediecuatorial, Quito, Ecuador: 364 pp.
Toulkeridis, T., 2013. Volcanes Activos Ecuador, Santa Rita, Quito, Ecuador: 152 pp.
Toulkeridis, T., Arroyo, C.R., Cruz D´Howitt. M., Debut, A., Vaca, A., Cumbal, L., Mato, F.,
Aguilera, E., 2016. Analysis of first ejected fine-grained material of reactivated Cotopaxi
volcano: Implications to activity & early warning, Volcanology & Seismology, submit.
Toulkeridis, T., Buchwaldt, R., Addison, A., 2007. When volcanoes threaten, scientists warn.
Geotimes, 52:36-39.
Toussaint, J.F., Restrepo, J.J., 1994. The Colombian Andes during Cretaceous times,
Cretaceous tectonics of the Andes, Vieweg & Teubner Verlag, 61-100.
Tran, K.T., Hiltunen, D.R., 2008. A comparison of shear wave velocity profiles from SASW,
MASW, and ReMi techniques, Geotechnical Earthquake Engineering & Soil Dynamics, ASCE
Geotechnical SP-181, Reston, VA.
UCSG, 2015. Seismic and geotechnical risk study for Rio Grande dam in compliance with
Ecuadorian code (Spanish: Estudio de Riesgo Sísmico y Geotécnico de la Presa Rio Grande de

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-14
Acuerdo a la Norma Ecuatoriana de la Construcción NEC-15), Technical Report, Facultad de
Ing. Inst. de Investigaciٕón y Desarrollo for Secretaría del Agua, Univ. Catolica de Guayaquil.
UNESCO World Heritage Sites, 2016. City of Quito, whc.unesco.org/ list/2, accessed 9/10/16.
UNICEF, 2016. Ecuador Earthquake: How to help children and families today, 4/21/16 article
by M. Sandler and M.G. Farley, unicefusa.org/stories/ecuador-earthquake-%E2%80%94-how-
help-children-and-families-today/30256, last accessed 6/9/16.
UNICEF, 2016. First airlift of UNICEF relief items reaches quake-hit Ecuador, 4/22/16 press
release, unicef.org/media/media_90992.html, last accessed 6/11/16.
Uniform Building Code, 1976. International Conference of Building Officials, Whittier, CA.
USGS, 2005. Volcanoes Glossary, United States Geologic Survey public domain,
commons.wikimedia.org/w/index.php?curid=3935253, last accessed 9/23/16.
USGS, 2015. Historic earthquakes off the coast of Ecuador, 1906 Jan. 31, US Geologic Survey,
earthquake.usgs.gov/earthquakes/world/events/1906_01_31.php, last accessed 5/24/16.
USGS, 2016a. Earthquake hazards program, United States Geologic Survey,
earthquake.usgs.gov/earthquakes/eventpage/us20005j32#executive, last accessed 6/12/16.
USGS, 2016b. M7.8 Muisne, EC, Earthq. Hazards Program, United States Geologic Survey,
earthquake.usgs.gov/earthquakes/eventpage/us20005j32#finite-fault, last accessed 5/18/16.
Vaca, V.A., Arroyo, R.C., Debut, A., Toulkeridis, T., 2016a. 2015 volcanic activity of
Cotopaxi ash data set, Imprenta ESPE, Sangolquí, EC, I:22.08-10.09.2015, 214 pp.
Vaca, V.A., Arroyo, R.C., Debut, A., Toulkeridis, T., 2016b. 2015 volcanic activity of
Cotopaxi ash data set, Imprenta ESPE, Sangolquí, EC, II:11.09-23.09.2015, 214 pp.
Vaca, V.A., Arroyo, R.C., Debut, A., Toulkeridis, T., 2016c. 2015 volcanic activity of
Cotopaxi ash data set, Imprenta ESPE, Sangolquí, EC, III:24.09-23.10.2015, 210 pp.
Vaca, V.A., Arroyo, R.C., Debut, A., Toulkeridis, T., 2016d. 2015 volcanic activity of
Cotopaxi ash data set, Imprenta ESPE, Sangolquí, EC, VI(11):24.10-19.11.2015, 220 pp.
Vallée, M. et al., 2013. Intense interface seismicity triggered by a shallow slow slip event in
Central Ecuador subduction zone, J. Geophys. Res. Solid Earth.
Vera, R. 2016. Geology of Ecuador, geologyofecuador.com/, last accessed 7/17/16.
Vera-Grunauer, X., 2014. Seismic Response of a Soft, High Plasticity, Diatomaceous Naturally
Cemented Clay Deposit, Doctoral Dissertation, University of California, Berkeley.
Vistazo Magazine, 2016. Dolor Infinito, Extraordinary Edition, April, 1168:82 pp.
Wathelet, M. et al., 2008. Array performances for ambient vibrations on a shallow structure
and consequences over Vs inversion, Seismology, 12:1-19.
Weiser-Woodward, A., May 25, 2016. Road Damage in Manabí, EERI Muisne Ecuador
Earthquake Clearinghouse, eqclearinghouse.org/2016-04-16-muisne/damage+in+manabi, last
accessed 9/9/16.
Werner, R., Hoernle, K., Barckhausen, U., Hauff, F., 2003. Geodynamic evolution of
Galápagos hot spot (Central East Pacific) in past 20 my: Constraints from morphology,
geochemistry, and magnetic anomalies. Geochemistry, Geophysics, Geosystems, 4(12).

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-15
White, S.M.M., Trenkamp R., Kellogg J.N., 2003. Recent crustal deformation and earthquake
cycle along Ecuador-Colombia subduction, Earth & Planetary Sc. Letters, 216(3):231-242.
Wikipedia, 2016. Provinces of Ecuador, en.wikipedia.org/ProvincesEcuador accessed 9/10/16.
Wolf, T., 1892. Geografía y geología del Ecuador, publicada por Orden del Supremo Gobierno
de la República por T. Wolf, Leipzig, Tipografía F.A. Brockhaus, reprint Alicante, Biblioteca
Virtual Miguel de Cervantes, 2006; cervantesvirtual.com/nd/ (USGS).
Wood, C., Ellis, T., Teague, D., Cox, B., 2014. Analyst I: Comprehensive analysis of UTexas1
surface wave dataset, ASCE GeoCongress: Geocharacterization & modeling for sustainability:
820-829.
World Atlas, 2016. Ecuador, worldatlas.com/webimage/samerica/ec.htm, accessed 9/10/16.
World Bank, 2015. Ecuador mortality rate, data.worldbank.org/locations=EC&start=2015,
accessed 9/10/16.
World Bank, 2016. data.worldbank.org/country/ecuador?view=chart, last accessed 6/30/16.
World Bank, 2016. Ecuador, worldbank.org/en/country/ecuador/overview, accessed 9/10/16.
Yepes, H., Audin, L., Alvarado, A., Beauval, C., Aguilar, J., Font, Y., Cotton, F., 2016. A new
view for the geodynamics of Ecuador: Implication in seismogenic source definition and
seismic hazard assessment, Tectonics, 35(5):1249-1279
Yépez F., Sanchez T., 2015. Design and construction considerations implemented in the
Ecuadorian construction code NEC2015”, Civil Engineers Association, Sigma, 12(30):8p.
Yépez, F., Yépez, O., 2016. Lessons learned from construction materials used in damaged
buildings after Mw7.8 Pedernales earthquake, April 2016, Technical Report, Universidad San
Francisco de Quito, USFQ, Ecuador.
Yoon, S., Rix, G., 2004. Combined active-passive surface wave measurements for near-surface
site characterization, Applications of Geophysics to Eng. & Env. Problems, 17:1556-1564.
Your Escape to Ecuador, 2016. The 4 Regions of Ecuador, yourescapetoecuador.com/the-four-
regions-of-ecuador/, accessed 9/10/16.
Youtube, 2016. Terrorifico! Camara capta terremoto en Ecuador, dated 4/16/16,
youtube.com/watch?v=Nz_05mT-Me0, released 5/10/16, last accessed 7/18/16.
Zywicki, D.J., 1999. Advanced signal processing methods applied to engineering analysis of
seismic surface waves. Ph.D. Dissertation, Georgia Inst. Technology, Atlanta, GA, 357 p.
Zywicki, D.J., 2007. The impact of seismic wavefield and source properties on ReMi estimates,
ASCE Geo-Denver: Innovative applications of geophysics in civil engineering, SP-164.

GEER-ATC Muisne, Ecuador, Earthquake 11-REFERENCES


Report Version 1 11-16
Blank Page
Blank Page
APPENDIX A

PRE-EARTHQUAKE GEOTECHNICAL INVESTIGATIONS

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-1
Blank Page
Figure A1. Pre-earthquake boring log B110, Manta Port (0°55'56''S, 80°43'13''W) (Ripalda, 2007).

Figure A2. Pre-earthquake boring log B111, Manta Port (0°55'50''S, 80°43'9''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-2
Figure A3. Pre-Earthquake boring log B116, Manta Port (0°55'50''S, 80°43'11''W) (Ripalda, 2007).

Figure A4. Pre-Earthquake boring log B117, Manta Port (0°55'25''S, 80°42'58''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-3
Figure A5. Pre-Earthquake boring log B118, Manta Port (0°55'40''S, 80°42'58''W) (Ripalda, 2007).

Figure A6. Pre-Earthquake boring log B120, Manta Port (0°55'37''S, 80°43'3''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-4
Figure A7. Pre-Earthquake boring log B123, Manta Port (0°55'40''S, 80°43'9''W) (Ripalda, 2007).

Figure A8. Pre-Earthquake boring log B124A, Manta Port (0°55'36''S, 80°43'2''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-5
Figure A9. Pre-Earthquake boring log B125, Manta Port (0°55'35''S, 80°42'58''W) (Ripalda, 2007).

Figure A10. Pre-Earthquake boring log B130, Manta Port (0°55'53''S, 80°43'16''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-6
Figure A11. Pre-Earthquake boring log B132, Manta Port (0°56'0''S, 80°43'20''W) (Ripalda, 2007).

Figure A12. Pre-Earthquake boring log B132B, Manta Port (0°55'58''S, 80°43'19''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-7
Figure A13. Pre-Earthquake boring log B134, Manta Port (0°56'4''S, 80°43'25''W) (Ripalda, 2007).

Figure A14. Pre-Earthquake boring log B135A, Manta Port (0°56'4''S, 80°43'28''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-8
Figure A15. Pre-Earthquake boring log B137, Manta Port (0°56'4''S, 80°43'20''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-9
Figure A16. Pre-Earthquake boring log B138, Manta Port (0°56'6''S, 80°43'23''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-10
Figure A17. Pre-Earthquake boring log B139, Manta Port (0°56'11''S, 80°43'24''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-11
Figure A18. Pre-Earthquake boring log B142, Manta Port (0°56'13''S, 80°43'21''W) (Ripalda, 2007).

Figure A19. Pre-Earthquake boring log B143, Manta Port (0°56'1''S, 80°43'14''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-12
Figure A20. Pre-Earthquake boring log B143A, Manta Port (Note: The exact location is unknown but
assumed to be near B143, as B143 had to be stopped at a shallow depth) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-13
Figure A21. Pre-Earthquake Test Pit TP-1, Manta Port (0°56'12''S, 80°43'31''W) (Ripalda, 2007).

Figure A22. Pre-Earthquake Test Pit TP-2, Manta Port (0°56'9''S, 80°43'30''W) (Ripalda, 2007).

Figure A23. Pre-Earthquake Test Pit TP-4, Manta Port (0°56'7''S, 80°43'28''W) (Ripalda, 2007).

Figure A24. Pre-Earthquake Test Pit TP-6, Manta Port (0°56'1''S, 80°43'19''W) (Ripalda, 2007).

Figure A25. Pre-Earthquake Test Pit TP-8, Manta Port (0°55'51''S, 80°43'12''W) (Ripalda, 2007).

Figure A26. Pre-Earthquake Test Pit TP-9, Manta Port (0°55'51''S, 80°43'12''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-14
Figure A27. Pre-Earthquake Downhole Seismic Vs testing at boring B-116, Manta Port (0°55'50''S,
80°43'11''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-15
Figure A28. Pre-Earthquake Downhole Seismic Vs testing at boring B-123, Manta Port (0°55'40''S,
80°43'9''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-16
Figure A29. Pre-Earthquake Downhole Seismic Vs testing at boring B-138, Manta Port (0°56'6''S,
80°43'23''W) (Ripalda, 2007).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-17
Figure A30. Soil Properties from Boring 1 (pre-earthquake), Mejia Bridge (0°59'24''S, 80°28'11''W)
(NYLIC ,2001).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-18
Figure A31. Soil Properties from Boring 2 (pre-earthquake), Mejia Bridge (0°59'23''S, 80°28'11''W)
(NYLIC ,2001).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-19
Figure A32. Soil profile at Los Caras Bridge (pre-earthquake, Courtesy of Ecuador’s Army Corps of
Engineers and Adolfo Caicedo).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-20
Figure A33. Top portion of boring log P1 (pre-earthquake), Las Chacras (Rio Chico) Bridge
(0°58'35''S, 80°25'21''W) (LUP, 2013).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-21
Figure A34. Bottom portion of boring log P1 (pre-earthquake), Las Chacras (Rio Chico) Bridge
(0°58'35''S, 80°25'21''W) (LUP, 2013).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-22
Figure A35. Top portion of boring log P2 (pre-earthquake), Las Chacras (Rio Chico) Bridge
(0°58'36''S, 80°25'23''W) (LUP, 2013).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-23
Figure A36. Bottom portion of boring log P2 (pre-earthquake), Las Chacras (Rio Chico) Bridge
(0°58'36''S, 80°25'23''W) (LUP, 2013).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-24
Figure A37. Boring log P1 (pre-earthquake), Briceño Bridge (0°30'59''S, 80°26'30''W) (Asesoria y
Estudios Tecnicos c. ltda, 2010).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-25
Figure A38. Boring log P2 (pre-earthquake), Briceño Bridge (0°30'56''S, 80°26'30''W) (Asesoria y
Estudios Tecnicos c. ltda, 2010).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-26
Figure A39. Boring log PPMCH-01 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'22''W) (TH,
2013) (Part 1 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-27
Figure A39. Boring log PPMCH-01 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'22''W) (TH,
2013) (Part 2 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-28
Figure A39. Boring log PPMCH-01 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'22''W) (TH,
2013) (Part 3 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-29
Figure A39. Boring log PPMCH-01 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'22''W) (TH,
2013) (Part 4 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-30
Figure A39. Boring log PPMCH-01 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'22''W) (TH,
2013) (Part 5 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-31
Figure A40. Boring log PPMCH-02 (pre-earthquake), Chone Dam (0°42'1''S, 79°59'23''W) (TH, 2013)
(Part 1 of 3).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-32
Figure A40. Boring log PPMCH-02 (pre-earthquake), Chone Dam (0°42'1''S, 79°59'23''W) (TH, 2013)
(Part 2 of 3).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-33
Figure A40. Boring log PPMCH-02 (pre-earthquake), Chone Dam (0°42'1''S, 79°59'23''W) (TH, 2013)
(Part 3 of 3).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-34
Figure A41. Boring log PPMCH-03 (pre-earthquake), Chone Dam (0°41'58''S, 79°59'21''W) (TH,
2013) (Part 1 of 4).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-35
Figure A41. Boring log PPMCH-03 (pre-earthquake), Chone Dam (0°41'58''S, 79°59'21''W) (TH,
2013) (Part 2 of 4).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-36
Figure A41. Boring log PPMCH-03 (pre-earthquake), Chone Dam (0°41'58''S, 79°59'21''W) (TH,
2013) (Part 3 of 4).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-37
Figure A41. Boring log PPMCH-03 (pre-earthquake), Chone Dam (0°41'58''S, 79°59'21''W) (TH,
2013) (Part 4 of 4).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-38
Figure A42. Boring log PPMCH-04 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'20''W) (TH,
2013) (Part 1 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-39
Figure A42. Boring log PPMCH-04 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'20''W) (TH,
2013) (Part 2 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-40
Figure A42. Boring log PPMCH-04 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'20''W) (TH,
2013) (Part 3 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-41
Figure A42. Boring log PPMCH-04 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'20''W) (TH,
2013) (Part 4 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-42
Figure A42. Boring log PPMCH-04 (pre-earthquake), Chone Dam (0°41'60''S, 79°59'20''W) (TH,
2013) (Part 5 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-43
Figure A43. Boring log PPMCH-05 (pre-earthquake), Chone Dam (0°41'59''S, 79°59'22''W) (TH,
2013) (Part 1 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-44
Figure A43. Boring log PPMCH-05 (pre-earthquake), Chone Dam (0°41'59''S, 79°59'22''W) (TH,
2013) (Part 2 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-45
Figure A43. Boring log PPMCH-05 (pre-earthquake), Chone Dam (0°41'59''S, 79°59'22''W) (TH,
2013) (Part 3 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-46
Figure A43. Boring log PPMCH-05 (pre-earthquake), Chone Dam (0°41'59''S, 79°59'22''W) (TH,
2013) (Part 4 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-47
Figure A43. Boring log PPMCH-05 (pre-earthquake), Chone Dam (0°41'59''S, 79°59'22''W) (TH,
2013) (Part 5 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-48
Figure A44. Soil properties from borings P1 – P5 (pre-earthquake), Chone Dam (P1:0°42'1''S,
79°59'21''W; P2:0°42'0''S, 79°59'21''W; P3: 0°41'59''S, 79°59'21''W; P4: 0°41'59''S, 79°59'23''W; P5:
0°42'1''S, 79°59'18''W ) (UCSG, 2015).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX A


Report Version 1 A-49
Blank Page
Blank Page
APPENDIX B

POST-EARTHQUAKE GEOTECHNICAL INVESTIGATIONS

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-1
Blank Page
Figure B1. Post-earthquake boring log P-1, Manta Port (0°55'58.8"S, 80°43'19.1994"W)
(GEOESTUDIOS, 2016) (Part 1 of 2).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-2
Figure B1. Post-earthquake boring log P-1, Manta Port (0°55'58.8"S, 80°43'19.1994"W)
(GEOESTUDIOS, 2016) (Part 2 of 2).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-3
Figure B2. Post-earthquake CPT, Manta Port (0°55'58.8"S, 80°43'19.1994"W) (Subterra Ingenieria
Geotecnica, 2016).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-4
Figure B3. Post-earthquake boring log P-1, Mejia Bridge (0°59'23.9994"S, 80°28'11.9994"W)
(GEOESTUDIOS, 2016) (Part 1 of 3).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-5
Figure B3. Post-earthquake boring log P-1, Mejia Bridge (0°59'23.9994"S, 80°28'11.9994"W)
(GEOESTUDIOS, 2016) (Part 2 of 3).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-6
Figure B3. Post-earthquake boring log P-1, Mejia Bridge (0°59'23.9994"S, 80°28'11.9994"W)
(GEOESTUDIOS, 2016) (Part 3 of 3).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-7
Figure B4. Post-earthquake CPT, Mejia Bridge (0°59'27.6"S, 80°28'11.9994"W) (Subterra Ingenieria
Geotecnica, 2016).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-8
Figure B5. Post-earthquake boring log P-1, APO1 Station (1°2'13.1994"S, 80°27'32.4"W)
(GEOESTUDIOS, 2016) (Part 1 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-9
Figure B5. Post-earthquake boring log P-1, APO1 Station (1°2'13.1994"S, 80°27'32.4"W)
(GEOESTUDIOS, 2016) (Part 2 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-10
Figure B5. Post-earthquake boring log P-1, APO1 Station (1°2'13.1994"S, 80°27'32.4"W)
(GEOESTUDIOS, 2016) (Part 3 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-11
Figure B5. Post-earthquake boring log P-1, APO1 Station (1°2'13.1994"S, 80°27'32.4"W)
(GEOESTUDIOS, 2016) (Part 4 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-12
Figure B5. Post-earthquake boring log P-1, APO1 Station (1°2'13.1994"S, 80°27'32.4"W)
(GEOESTUDIOS, 2016) (Part 5 of 5).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-13
Figure B6. Post-earthquake boring log P-1, Mobil (0°56'52.7994"S, 80°43'15.6"W) (GEOESTUDIOS,
2016) (Part 1 of 2).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-14
Figure B6. Post-earthquake boring log P-1, Mobil (0°56'52.7994"S, 80°43'15.6"W) (GEOESTUDIOS,
2016) (Part 2 of 2).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-15
Figure B7. Post-earthquake boring log P-2, Mobil (0°56'49.2"S, 80°43'8.3994"W) (GEOESTUDIOS,
2016) (Part 1 of 2).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-16
Figure B7. Post-earthquake boring log P-2, Mobil (0°56'49.2"S, 80°43'8.3994"W) (GEOESTUDIOS,
2016) (Part 2 of 2).

GEER-ATC Muisne, Ecuador, Earthquake APPENDIX B


Report Version 1 B-17

You might also like