Dumont Article 2006

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Geomorphology 74 (2006) 100 – 123

www.elsevier.com/locate/geomorph

Fan beheading and drainage diversion as evidence of a 3200-2800


BP earthquake event in the Esmeraldas-Tumaco seismic zone:
A case study for the effects of great subduction earthquakes
J.F. Dumont a,*, E. Santana b, F. Valdez c, J.P. Tihay d, P. Usselmann e,
D. Iturralde f, E. Navarette g
a
IRD, UMR 6526, Observatoire Océanographique, 06235 Villefranche sur mer, France
b
INOCAR, Base Naval Sur, Av. de la Marina, Guayaquil, Ecuador
c
IRD, UR 092, Instituto Nacional de Patrimonio Cultural , Quito, Ecuador
d
UPPA, Pau, France
e
CNRS-UMR 6012, Maison de la géographie, 34090 Montpellier, France
f
Universidad de Guayaquil, Guayaquil, Ecuador
g
ESPOL, Campus La Prosperina, Guayaquil, Ecuador
Received 7 October 2004; received in revised form 25 June 2005; accepted 27 July 2005
Available online 22 September 2005

Abstract

The San Lorenzo area belongs to the Esmeraldas–Tumaco seismic zone where some of the strongest earthquakes of South
America occurred during the 20th century. This paper provides evidence for a succession of geomorphic changes characterized
by the disruption of the Quaternary drainage network and the reshaping of the Cayapas–Santiago estuary. The rise of the La
Boca uplift bordered by the La Boca and San Lorenzo faults is responsible for the southward diversion of the Palabi, Tululbi,
Bogotá and Carolina rivers toward the Santiago and Cayapas rivers. The increase of the discharge directed to the Cayapas River
generated the change of the channel pattern downstream from the confluence, and the avulsion of a new estuary through the
coastal plain. According to the dating of beach ridges the avulsion occurred in the period 3200–2800 BP. This period
corresponds also to a faster accretion of the beach ridge margin, interpreted as a response to a small uplift of the shore. The
coherency of the three morphologic evidences—diversion of drainage network, avulsion and increase of coastal accretion—
suggest a unique morphotectonic event, in relation with the activity of the Esmaraldas–Tumaco seismic zone. The opening of a
direct communication through the mangrove margin may have brought favorable conditions for the development of the La
Tolita archaeological site after 3000 BP.
D 2005 Elsevier B.V. All rights reserved.

Keywords: Active tectonics; Drainage diversion; Drainage pattern; Beach ridges; Ecuador

* Corresponding author.
E-mail address: dumont@geoazur.obs-vlfr.fr (J.F. Dumont).

0169-555X/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.geomorph.2005.07.011
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 101

1. Introduction 1981), the largest being the 1906 Tumaco earthquake


(Mw 8.8), associated with a rupture zone of 500 km
Active tectonics of coastal areas involves various (Collot et al., 2004; Herd et al., 1981). Co-seismic
types of natural hazard, earthquake, tsunamis and deformations, either uplift (Barrientos and Plafker,
modification of coastal morphology. Paleoseismology 1992; Ortlieb et al., 1996), or subsidence (González
is based on several kinds of evidences, among them et al., 2002; Herd et al., 1981), are reported in asso-
the deformation of land surface due to fault motion ciation with subduction earthquakes along the Pacific
(McCalpin, 1996; Mörner, 2003). Active tectonics coast of South America. These co-seismic and inter-
involves a wider approach, in order to detect active seismic deformations depend on the distance to the
deformation and document possible paleo-earthquake seismic zone (Barrientos, 1996; Barrientos and Plaf-
events (Keller and Pinter, 2002). The North Andean ker, 1992; Carver and McCalpin, 1996; Fitch and
coast has produced four great earthquakes with mag- Scholz, 1971). A null elevation change axis frequently
nitude Mw N7.7 during the last century (Fig. 1A) located near the coastline separates areas of alternat-
(Beck et al., 1998; Gutscher et al., 1999; Herd et al., ing uplift and subsidence during seismic and inter-

80°W
Esmeraldas-Tumaco seismic zone 197
Fig. 2
1906 (Mw 8.8)
*+
9
06

Plate Tumaco Plate


1942 (Mw 7.8) 58
19

convergence 19 convergence
1958 (Mw 7.7) 1958
ch

1906 Colombia
tren

1979 (Mw 8.2)


42

* Esmeraldas
19

sin
tion

Carchi
Ba

Nazca Plate
ón
duc

rb

Ecuador
Bo

1942
Sub

ra

shear
+
lle

ra

rdi

Quito
Co

dille
Mega

Carnegie
Cor
s

Ridge
a
l ta

rac
Coas

Ca

80°W 70°W
Carribean
-

CARACAS
-2000 m i l
Plate 10°N
aqu
y
ua
North
s

Guayaquil
ea ca

Andean

Andean
G
sh ara

Block
r
ga - C

Nazca
Me uil

Plate
aq
ay

Fig.1A
Gu

0° Gulf of
South American Guayaquil
Talara Arc

A
00m

Plate
- 10

GUAYAQUIL
100 km
B
Fig. 1. Geodynamic sketch of the Ecuadorian active margin. Rupture zones and magnitudes of the 1906, 1942, 1958 and 1979 earthquakes come
from Collot et al., 2004.
102 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

seismic deformations (Barrientos, 1996; Fitch and et al., 2002), defined in reference to the great 1906
Scholz, 1971). Up to 8 m uplift and 6 m subsidence earthquake, one of the strongest earthquake regis-
have been observed in relation to the 1946 magnitude tered on earth (Mw 8.8). The rupture zone of the
8.2 Nankaido earthquake in Japan, and surface defor- 1906 earthquake has been reactivated by 3 strong
mation has been continuing 3 years after the earth- earthquakes in 1942 (Mw 7.9), 1958 (Mw 7.8) and
quake (Fitch and Scholz, 1971). Herd et al. (1981) 1979 (Mw 7.9) (Collot et al., 2004; Herd et al.,
report co-seismic subsidence of up to 1.6 m in relation 1981; Kelleher, 1972; Swenson and Beck, 1996),
to the 1979 magnitude 7.8 Tumaco earthquake. and strong earthquakes also probably occurred in
This study supports the hypothesis that co-seismic the same area in 1868 and 1875 (Scheu, 1911).
deformation over a subduction zone may also produce The 1979 earthquake has been accompanied by a
change of the drainage network and/or adjustment of coastal subsidence of up to 1.6 m, a ground failure
the drainage pattern that can be useful for paleoseis- extending up to 60 km inland, and co-seismic uplift
mology of extreme events. In flat areas and wet is reported offshore, on the edge of the continental
climate, light vertical motion of the surface morphol- shelf (Herd et al., 1981).
ogy can generate striking changes of the drainage The coastal margin of central Ecuador is character-
pattern. This approach has been previously used to ized by a steady uplift during the Quaternary giving
study active surface deformation in wetland and allu- flight of marine terraces up to an elevation of 300 m
vial plains of the Amazonian foreland basins (Pedoja, 2003). This uplift is related to the subduction
(Dumont, 1991, 1993; Dumont and Fournier, 1994; of the Carnegie Ridge (Gutscher et al., 1999; Pedoja
Schumm et al., 2000). There, low range vertical defor- et al., 2001). The segment of uplifting coast stops to
mations resulted in river diversion and channel pattern the north on the NW–SE trending Yanayaca fault (Fig.
change along small rivers, and avulsion of the main 2) (CERESIS, 1985; Pedoja, 2003). From Las Peñas
streams. Similar climate and drainage conditions exist to the Manglar Cap the Bay of Ancón de Sardinas
in the Esmeraldas–Tumaco coastal margin of North- (Fig. 2) constitutes a wide delta formed between the
ern Ecuador and Colombia. outlet of the Cayapas–Santiago and Mataje rivers,
located on the southern and northern borders of the
delta, respectively. The delta area constitutes a 15-km
2. Geological background wide margin of mangrove, beach ridges and tide
channels, bounded sharply inland by a continuous
The Nazca and South American plates converge 10–15 m cliff. This cliff also represents the limit of
along the South American active margin at a rate of 7– the mid Holocene post-glacial transgression (Tihay
8 cm/yr with a nearly W–E trend (De Mets et al., and Usselmann, 1995).
1990; Kellogg and Vega, 1995). From the Gulf of The Pleistocene deposits of the San Lorenzo area
Guayaquil (south) to the Ecuador–Colombian border are represented by the Cachabi (also called Canoa)
(north) the obliquity of the subduction zone with the and San Tadeo formations (Fig. 3) (Baldock, 1982;
plate motion increases up to 30–50 degrees (Fig. 1A) CODIGEM, 1993). The regional extension of these
(Ego et al., 1996) . This oblique subduction favors the formations describes a large fan along the lower
northeastward escape of the North Andean block at a slopes of the Andean Cordillera, down to the coast.
rate of about 1 cm/yr along the Guayaquil–Caracas The uppermost San Tadeo Formation crops out along
megashear (Dumont et al., 2005; Ego et al., 1996) the northern slopes of the fan. It includes pyroclastics,
(Fig. 1B). The northward motion of the Andean Block volcanic conglomerate, laharic and mudflow material,
goes along with a roughly N–S trending extension in forming sheets or fluvial terraces entrenched in the
the Ecuadorian coastal margin, involving NE–SW Cachabi Formation (Baldock, 1982). The Cachabi
dextral and NW–SE sinestral transtension faults Formation represents a continental fan made of
(Alvarado, 1998; Deniaud, 2000; Dumont et al., sand, mud, tuff, gravel and conglomerate. The mate-
1997, 2005; Santana and Dumont, 2002). rial is deeply weathered in reddish-yellow clay, with
The San Lorenzo area is located landward of the only ghosts of siliceous pebbles showing the original
Esmeraldas–Tumaco seismic zone (Fig. 1A) (Collot facies.
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 103

Fig. 2. Geologic sketch of the area of San Lorenzo, localised on Fig. 1. Numbers 1, 2 and 3 represent the different drainage areas described in the
text.

The basement of the Cachabi and San Tadeo for- 1999; Deniaud, 2000; Evans and Whittaker, 1982).
mations belongs to the Borbón Formation of the The depocenter of the north Borbón Basin is located
Borbón Basin, that was active since the Middle below the Ancón de Sardinas Bay, near the estuary of
Eocene to the Upper Pliocene (Aalto and Miller, the Cayapas–Santiago river and the city of San Lor-
104 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

Fig. 3. Synthetic stratigraphic section showing the relative position of the different units. No scale.

enzo (Deniaud, 2000). The location of the delta on a geologic (CODIGEM, 1993) and 1 : 200,000e geomor-
structural low underlines the long term structural his- phologic map (PRONAREG-ORSTOM, 1984a,b,c).
tory of this area, as it is generally observed for the This overview gives evidence of a discontinuous net-
lower valleys and estuaries of large rivers (Potter, work suggesting diversion and re-routing apart from a
1978). Folding and emergence of the south Borbón previous straight down slope drainage network (Figs.
Basin post-dates the late Pliocene (Evans and Whit- 2 and 3). A drainage network flowing against or very
taker, 1982). obliquely to the regional slope may be suspected of
tectonic control (Booth-Rea et al., in press; Jackson et
al., 1996; Keller and Pinter, 2002; Schumm et al.,
3. Methods 2000) as it is observed in some parts of these maps.
However, Bishop (1995) pointed out that most of
We apply a three-step method comprising: 1–the drainage diversion–or beheading–and re-routing may
global analysis of the drainage network, 2–the be explained as simple capture, and each case must be
detailed analysis of drainage re-routing, apparent analyzed carefully before stating active tectonics. Pas-
channel pattern anomalies, and avulsion, and 3–the sive drainage capture is generally made by a more
search for complementary elements, analyzing the incised or active drainage segment reaching a less
coastal plain and the other evidences of active fault active one upstream. But active tectonics (fault, flex-
tectonics from the same area. ure) can open drainage by-pass toward less active
The analysis of the regional drainage network con- drainage segments (Bishop, 1995; Bowler and Har-
stitutes customarily the first step of the search for ford, 1966; Dumont et al., 2005; Jackson et al., 1996).
evidences of active tectonics (see Schumm et al., Comparison with examples previously described in
2000; Keller and Pinter, 2002 for a review on the the literature is important, and the river re-routing in
question). The first step considers the main trend of La Boca described in this paper is comparable to the
the present drainage network from the Western classical example of a river blocked and re-directed
Andean Cordillera to the coast on 1 : 250,000e topo- against a new or reactivated fault, as described by
graphic maps (IGM, 1991, 1992a,b), 1 : 1000,000e Bowler and Harford (1966).
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 105

The second step comprises the detailed analysis of River, and the first two converge at 10 km to the north
some characteristic points of the drainage diversions (Figs. 2 and 6). This upstream convergence of the
that disrupt the original fan network, at the head of the drainage describes a fan network called the El Placer
fan, in the middle part, and finally in the coastal plain. fan (Winckell and Zebrowski, 1997). In the down-
Channel slope, channel trend change, as well as dry stream part of area 2 (Fig. 2) the drainage network of
valley, are important elements to analyze and interpret the El Placer fan is sharply re-routed to the SW, the
the reshaping of the drainage network. Downstream Palabi, Tululbi, Bogotá and Carolina rivers joining
the drainage network enters the delta area, showing together the Santiago River. Along the NE margin
combination of anastomosed and meandering pat- of the El Placer fan the Mataje River flows straight
terns. Different channel patterns with the same slope down slope, reaching the coast along the structural
imply changing hydrologic characteristics such as border of the Bay of Ancón de Sardinas.
discharge and sediment transport (Schumm et al., Area 3 (Fig. 2) covers a 30-km wide lowland
2000), that should be a key for the understanding of margin (0–50 m), including the San Lorenzo upland
drainage pattern in the coastal plain. and the coastal plain. The San Lorenzo upland is
The third step considers complementary elements characterized by a poor drainage network in the
that belong to the geographic and structural context. central part, with the Cayapas–Santiago and Mataje
In the coastal plain we analyze the space–time accre- rivers located at the southern and northern borders of
tion of the beach ridges in order to date the avulsion the area, respectively. The 15-km wide and 50-km
and the opening of the present Cayapas–Santiago long coastal plain includes mangrove, tidal channels,
estuary. Also space–time diagram of the beach ridge salted marshes and beach ridges, and the estuaries of
accretion will specify the coastal plain evolution. the Cayapas–Santiago and Mataje rivers. A striped
Finally the evidences of active fault motion are ana- pattern of beach ridges is observed in the southern
lyzed in the context of the morphological changes part, disappearing progressively to the north below
observed. mangrove and tidal channels. The Cayapas and San-
tiago rivers drain the Western Andean Cordillera
south of the Mira River catchment area. These rivers
4. The drainage network join together before to cross the delta through the
common Cayapas–Santiago estuary. A total of four
Between the Carchi area in the Western Andean large tide channels cross the coastal plain, from
Cordillera and the Bay of Ancón de Sardinas (Fig. 2) south to north: the Cayapas–Santiago, Boca de
the drainage network presents three parts (IGM, Limones, Bolivar and Mataje channels (Fig. 4). But
1992a,b). The area 1 (Fig. 2) extends from the Carchi only the Cayapas–Santiago and Mataje channels
area (about 2000 m) to Alto Tambo (500 m), and is located, respectively, in the SW and the NE are the
drained along by the Mira River, one of the main outlets of important rivers. The other ones constitute
rivers of northern Andes. The Mira River valley is the outlet of very small rivers without relation with
entrenched about 1000 m in the upper slopes of the the size of the channel (Fig. 4). Near San Lorenzo
Cordillera, and about 200–300 m in the Alto Tambo the Bolivar channel is connected landward with the
area. At Alto Tambo the trend of the Mira River turns Estero El Salto (Fig. 4), characterized by a typical ria
to the north, joining the catchment area of the San morphology carved in the San Lorenzo upland, and
Juan River, and reaching the coast in the Manglar Cap interpreted, as it is generally the case for such pat-
in Colombia (Fig. 2). tern, either for eustatic or tectonic effect, or both,
Area 2 (Fig. 2) extends over 40 km downslope suggesting a drowned Holocene valley.
from the Alto Tambo area (500–700 m) to a NE–SW
trending line joining Tululbi to Concepcion (20–50 4.1. Reconstruction of the Pleistocene fan drainage
m). The drainage network of area 2 comprises from
north to south the Mataje, Palabi, Tululbi, Bogotá and The different segments of the present drainage net-
Carolina rivers. The trend of the three last rivers work are reconstructed on Fig. 5, describing a large fan
converge upstream in Alto Tambo, close to the Mira with the Mira River as the fan feeder at the head, the El
106 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

Fault Ancón
COLOMBIA
Estimated 2800-3200 BP Mangrove
beach ridge line Ma
taj
Beach ridges eF
Upland over 20 m au
lt
Pleistocene fan Floodplain area Rio
Upland below 20 m Ma
Upper Cayapas terrace Limit of recent taje

Boliv
beach ridges
+ Bank of dead oyster pattern Q
ECUADOR

ar cha
N 7
σ3 Bay of Ancón
B1

LF
4 6
1

nnel
Nm

2
de Sardinas

F
5

WS
3

ESL
6
4
7 2 3 1
σ2 5 0 5 10 15 t,s Z' Fig. 12
σ1

San Estero
N Lorenzo
Li X4
σ2
B2 H m Bo
on c
es a d
M
N
O
Fig. 11
El Salto

Valdez ch e B1
Ca an P
X3
ya n el
σ1
σ3
p + Fig. 9
es as-
tu Sa
ar nt
y ia L La Boca
go X2 Uplift
La Tolita K
F

G
SL

J
F I
Y2 Z X1
C
D

E
A
B
Las Peñas
Y1 0 5km
A
A: 658-469 BP * G: 2781-2340 BP * M: 3200-2780 BP +
B: 771-539 BP * H: 460-0 BP + N: 5050-4860 BP +
C: 3471-2983 BP * I: 5935-5581 BP * O: 5450-5130 BP +
D: 3251-2854 BP * J: 5938-5600 BP * P: 6880-6660 BP +
E: 3260-2830 BP * K: 4132-3686 BP * Q: 3840-3580 BP +
F: 1707-1299 BP * L: 4230-3806 BP *
* from Tihay and Usselmann (1995) + This study
Ages from the beach ridges of the coastal plain
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 107

Placer fan as the main part, and the large tidal channels security) did not allow field work in this area. The
of the coastal plain as former estuaries. This drainage northward turn of the Mira River in Alto Tambo, from
network reconstruction is superimposed over the Pleis- NW–SE to N–S, is considered as a diversion because
tocene continental deposits of the Cachabi and San it drives the fan feeder river outside of the fan, into the
Tadeo formations (CODIGEM, 1993; PRONAREG- adjacent catchment area of the San Juan River. During
ORSTOM, 1984a,b,c; Winckell and Zebrowski, this process the Mira River passes over the structural
1997), and constitutes in our interpretation the last border of the fan, represented by the escarpment of the
evidence of the activity of this fan. Mataje fault. We will describe successively the trend
The geographic and geologic setting shows that the and slopes of the river segments involved, and the
El Placer fan is structurally enclosed in a system of structural fault pattern of the area before to interpret
NW–SE trending faults (Figs. 2 and 5). The south- the process of the diversion.
western border presents a step-like fault arrangement In Alto Tambo (Fig. 6, point A) the trend of the
from Las Peñas to Alto Tambo, and the northeastern Mira River changes from N558W F 158 (upstream) to
border is continuously bounded by the Mataje fault North F 108 (downstream), with a short intermediate
(Fig. 2). The Mataje fault separates the hilly upland on segment trending NE–SW in the turn. The trend of the
the Colombian side from drowned beach ridges and Bogotá River at the head of the El Placer fan, and the
mangrove on the Ecuador side, evidencing the activity Mira River upstream from Alto Tambo (La Tigrera)
of the fault during the Holocene. Another NW–SE are lined up, the head of the fan standing about 200 m
trending fault controls the position of the Mira River higher than the bottom of the Mira River (Fig. 6, point
upstream from the head of the fan (CODIGEM, A). About 10 km downstream from Alto Tambo the
1993). Mira River presents another turn to the NE. (Fig. 6,
The present segmentation of the drainage network point B). The Mira River valley upstream from the
suggests the disruption of a large fan network in three turn is lined up with the head drainage line of the
parts: at the head of the fan in Alto Tambo (between Mataje River, also located about 200 m higher. Down-
areas 1 and 2, Figs. 2 and 5) and in the middle part stream from Alto Tambo the Mira River follows the
(between areas 2 and 3, Figs. 2 and 5). The Mira River border of the El Placer fan, between the radial–locally
(the fan feeder) has been re-routed to the north, join- parallel–drainage network of the fan to the west and
ing presently the catchment area of the San Juan River the dendritic drainage network of the Andean slopes
in Colombia. Between areas 2 and 3 the drainage of to the east.
the El Placer fan is collected and re-routed to the SW The slope of the Mira River valleys is about 1.40%
toward the Santiago River (Figs. 2 and 5). The present (1.2%–1.5%) through the Western Cordillera
study analyses with more detail the morphology, drai- upstream from Alto Tambo, and drops to 1.2%
nage diversion that occurred at the lower part of the (1.1%–1.3%) downstream. However just before the
fan, in the areas 2 and 3. northward turn the Mira River valley widens, forming
the La Tigrera alluvial plain with a slope lower than
4.2 Diversion of the Mira River in Alto Tambo 0.6%. This alluvial fill implies a restraining of the
sedimentary transport just before the short NE–SW
The diversion of the Mira River at the head of the trending segment located at the beginning of the bend
fan is analyzed on maps (CODIGEM, 1993; IGM, (Fig. 6, point A). A similar scheme is repeated down-
1992a,b), because local conditions (access facility and stream in point B, before the turn to the NE. At the

Fig. 4. Morphostructural sketch of the Cayapas–Santiago deltaic area. A: Letters in squares refer to calibrated radiocarbon dating marked below
the figure. B1: Stereonet of the fault plane data from the San Lorenzo Fault, located on part A of the figure. Lower hemisphere projection of
Wulf net. Letters j1, j2 and j3, respectively, represent the maximum, intermediate and minimum stress axis calculated by the Carey method
(1979). Arrows represent slickensides. The histogram represents the deviation between the observed and calculated data. Fault plane data
(azimuth-dip-pitch of the slickenside-motion): 1: N045-75N-60E-norm; 2: N055-85N-60N-norm; 3: N055-85-70N-norm; 4: N040-90-40N-
norm; 5: N090-80N-82W-norm; 6: N065-80W-80N-norm; 7: N080-85S-80W-norm. B2: State of stress obtained from the inversion of focal
solutions of subduction earthquakes, for the area of central and northern Ecuador, according to Ego et al. (1996).
108 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

Post-fan faults Tumaco

Fan related faults

Tambo Alto fan Manglar


Cap
Post fan
drainage

Abandoned
drainage M
at
aj
Observed drainage e G
Fa ua
ul lp
t i(
M
Ancón ira
)
R
iv
er
Bay of Ancón
de Sardinas

San Lorenzo San Juan River


La Tola

3
Ca
ya stua

Pa
pa ry

Mat
e

Bi g

Tu labi
s-S

Las Peñas lu
aje River
ua

R.
ra lb
an

lR iR
ti

. .
ag

Ya
n

2
ay
o

ac Borbon
aF
Bo

au
go

lt
Old

ta
R
S

.
anti

Ca
ro
ago
r

lin
ve

aR
Ri

. Alto Tambo
Riv
s
pa

er
ya

Mi
Ca

ra
Riv
er

0 20 km
1
Fig. 5. Reconstruction of the drainage network of the Pleistocene Alto Tambo fan, combining the drainage areas 1, 2 and 3 of Fig. 2.

head of the El Placer fan the slope of the drainage the borders of the El Placer fan. The NE–SW trending
lines (Tululbi, Bogotá rivers) is higher, about 5%–7% faults appear on both sides of the Mira River valley,
(Fig. 6, point C). one segment to the north and two 4-km spaced seg-
The geological map (CODIGEM, 1993) shows that ments to the south (Fig. 6). These north and south
the El Placer fan is located at the intersection of two segments are about lined up (Fig. 6), the continuous
directions of faults trending NE–SW and NW–SE, fault trace crossing the Mira River valley along the
respectively (Figs. 2 and 6). Upstream from Alto short NE–SW deviation. The NE–SW trending faults
Tambo the Mira River valley follows a NW–SE trend- constitute the northeastward extension of the Rio
ing fault, which is parallel to the faults observed on Canandé fault (Alvarado, 1998; Eguez et al., 2003)
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 109

Mira R.
Ma
taj
Tu

eR
lu lbi

.
100 R.

500
Bo
go
tá B
R.
0
100

500
Ca
ch
ab
iR
C A
.
La Tigrera
Caro
lina R
.
0 Alto Tambo
50 El Placer Mir
aR
ive
r
0
100
e

10
on
r

lt Z

00
e

00
go Riv

au

20
éF
nd
Santia

na

N
Ca

00
10
10 km
2000

Fig. 6. Morphologic sketch of the head part of the El Placer fan. Thick lines are faults from the geological map (CODIGEM, 1993), and thick
dotted lines are extrapolated faults with high probability, and thin dotted lines subdued possible faults. Letters refer to commentary in the text.

(Fig. 6), suspected of right-lateral Quaternary motion corresponds to the fault motion since the entrench-
on the basis of observation on radar images (Alvar- ment of the Mira River, that is since no more than
ado, 1998). The North Canandé fault is defined here approximately the middle Pleistocene. This gives a
as the extension of the Canandé fault in the Alto rough estimation of a slip rate of 1–2 mm/year.
Tambo area. The NE–SW river segments north of It is difficult to apply the same analysis to the other
the Canandé fault is interpreted as another segment NE–SW trending segment downstream from point B
of the Canandé fault zone. (Fig. 6), because it is located in an area of relatively
The Mira River valley presents a right hand offset low topography suggesting weak tectonic control and
of 1.5–2 km at the crossing point with the Canandé possibly easier natural diversions and adjustment
fault. The occurrence of the El Tigrera alluvial plain (Dumont et al., 2005; Huang, 1993). In those cases
supports the hypothesis of restraining drainage along the dextral apparent offset can be far higher than the
the fault segment, and thus the recent motion of the tectonic offset along the fault. However the tectonic
fault. This allows speculation that the river offset offset along the Canandé fault and the parallel fault to
110 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

the north are not enough to explain the diversion of case the capture of the Mira River occurs along drai-
the Mira River outside the structural limits of the El nage lines located at the structural border of the fan
Placer fan. We also need to consider the N–S trend- with the Andean Cordillera.
ing segments driving the Mira River to Colombia. Since the diversion occurred the Mira River has
The geological map does not show N–S trending been entrenched about 200 m relative to the head of
faults in the studied area, but such faults are frequent the El Place fan. This suggests that the diversion
in the southwestern part of the Borbón Basin, along probably occurred before the Holocene. A rough cal-
the Esmeraldas River valley (Eguez et al., 2003), and culation of a minimum age can be made considering
near the coast (Santana et al., 2001; Witt, 2001). In the entrenchment of the Pastaza River inside its fan on
particular the two more recent tectonic events the other side of the Andes, under similar climate. The
observed in the Pliocene deposits are successively calculation using the 5–6.7 mm/year 1 entrenchment
E–W and N–S extensions (Santana et al., 2001; Witt, of the Pastaza River (Bès de Berc et al., 2005) as a
2001), but only the last one affects the Holocene maximum rate gives a minimum of about 40,000
(Santana and Dumont, 2002) as it will be shown years for the entrenchment of the Mira River, since
later in the San Lorenzo area. We interpret here the the abandonment of the fan drainage network.
NS trending river segments as controlled by faults More probably, as suggested before, a middle
formed or reactivated during the early E–W exten- Pleistocene age can be suggested for the beginning
sion event of the Quaternary, before the Holocene. of the diversion at the head of the fan.
NE–SW as well as NW–SE trending faults are able
to have been moved during this event. This exten- 4.3 Southward deflection of fan drainage
sion tectonic event is able to have reduced the
activity of the fan. The result would be the opening The Palabi, Tululbi and Bogotá rivers are sharply
of new drainage lines toward the north along fault re-routed to the southwest against the La Boca upland
segments linking new or pre-existing drainage lines. in the downslope part of area 2 (Fig. 2). The topogra-
The entrenchment of the Mira River along the border phy of the La Boca area is only 20–30 m over the
structures of the fan occurred during this period. The bottom of the valleys, but defines a continuous
more recent and still active N–S extension did not watershed between the middle and lower parts of the
drastically change this pattern of morphostructures. El Placer fan (areas 2 and 3, Fig. 2). We studied in
This recent event is characterized by transtension particular the turn of the Tululbi and Bogotá rivers
motion, and dextral motion along the NE–SW trend- near Carondelet (Fig. 7) because of the relatively easy
ing faults (Santana and Dumont, 2002). It may be access and its representativeness as drainage diversion.
hypothesized that this recent motion is more pre- North of Carondelet the Tululbi River flows in an
cisely responsible for the restraining flow where opposite trend relative to the regional slope, as it has
the Canandé fault cuts the Mira River valley at been re-routed against the La Boca upland 5 km to the
point A, resulting in the formation of alluvial fill north (Fig. 2). The Bogotá River reaches the La Boca
like that of La Tigrera. upland near Carondelet, flowing inside a 30–40 m
According to Bishop’s (1995) classification of river deep and 0.15% sloping valley incised in the Cachabi
diversion the process observed here involves the pre- Formation. It merges with the Tululbi River at an
servation of drainage lines and transfer of drainage elevation of 14 m against the La Boca upland. Down-
areas between adjacent catchments. This is a btop– stream from the turn, the slope is only 0.03%, with a
downQ process, the tectonism at the point of river meandering pattern at the surface of a wide floodplain
capture determining the rearrangement of the catch- that extends widely to the SE. The border of the La
ment area downstream. According to Bishop (1995), Boca upland is 15–20 m over this floodplain. This
rejuvenation head may appear at or above the elbow border is interpreted as a fault that created the mor-
of capture, depending on the original height difference phology that re-routed the Bogotá River to the south
between the bed of the two rivers involved in the together with the Tululbi River. A key element is the
rearrangement and the change in discharge for the observation of a dry valley across the La Boca upland,
stream receiving the diverted flow. In the present in continuation of the Bogotá River upstream from
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 111

Tu
lulb
21 m .

iR
.
.
R 50m
pi .
Zas
48m
.

35m
. Slo
pe La Boca
La Boca 0.1
4%
uplift 28m . .14 m
Bo Carondelet
lt go
Fau ta
R.
45m
. ca
a Bo %
03
L

Sl
0. op
ope e
0.
Sl 15
%
Railway

2 km

Fig. 7. Morphologic map of the Carondelet area localized on Fig. 2. See commentary in the text.

Carondelet. A slope of 0.14% is observed across the River at La Boca suggests that the change has been
upland, comparable to the 0.15% slope of the Bogotá relatively sudden, as no intermediate morphology
river valley upstream from La Boca. The difference of attests to a progressive or stepped evolution. This
elevation between the bottom of the Bogotá River does not mean that the 15-m vertical offset of the
channel and the dry valley at La Boca reach 15 m, La Boca upland is the result of one tectonic event,
that is interpreted as the vertical relative offset of the but that one event has been enough to make the
La Boca Fault. diversion definitive. The Bogotá River is entrenched
To summarize we interpret the turn of the Bogotá 2–3 m in the floodplain in the Carondelet area,
River at La Boca as the result of the motion of the giving an estimation of the minimum vertical offset
La Boca Fault, that rises the upland and blocked the necessary to re-route the river. The morphostructural
drainage (Fig. 7). A similar situation explains the re- and drainage pattern at La Boca is comparable, on a
route of the Tululbi River to the south, along a fault smaller scale, to the Australian example of the dis-
segment parallel to the La Boca Fault and lined up ruption of the Murray River against the uplift of the
with regional morphostructures (Fig. 2). The mor- Cadell Block in Australia (Bowler and Harford,
phologic pattern at the turning point of the Bogotá 1966).
112 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

5. Channel pattern of the Cayapas–Santiago mosed pattern that appears downstream from the con-
confluence fluence between the Cayapas and Santiago Rivers,
and continue as a wide estuary through the outer
Between the confluence of the Cayapas and San- margin of the delta downstream from point 3. The
tiago Rivers and the coast we observe crossed chan- Los Atajos River (segment 2–5) is lined up with the
nels of different patterns, anastomosed and Cayapas River upstream from point 1. The Los Atajos
meandering (Fig. 8). The SPOT image (Fig. 8) River is presently a secondary channel with lesser
shows the main elements of the drainage network, discharge and limited navigation capacity with respect
numbered 1 to 5. Downstream from the confluence to the Cayapas–Santiago estuary.
of the Cayapas and Santiago rivers (point 1) the main A meandering channel pattern appears upstream
active channel runs presently through points 2, 3 and from point 1 and between points 2 and 5. The char-
4. The segment 1–3 is characterized by an anasto- acteristics are similar with a meander wavelength of

5
Ca
4 ya
pa
s
es -San
tua tia
ry go R.
3
jos
Ata

G: 2781-2340BP
Los

A A'

2
D: 3251-2854 BP
E: 3260-2830 BP

San
1 tiag
oR
.
R.
as
yap

5 km
Ca

Fig. 8. Extract of the SPOT image 638-347 of 22/03/94 showing the Cayapas–Santiago estuary through the deltaic shoreland. See localization
on Fig. 2. Points 1 to 5 refer to the different segments of channel pattern described in the text. A and A’ represent the expected extension of
floods along the upper part of the Cayapas–Santiago channel.
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 113

3–5 km, a sinuosity of about 1.8 (River channel Cayapas–Santiago estuary presents a similar pattern
length/river valley length) and channel width of to the Boca de Limones and Bolivar tidal channels,
150–250 m. The ratio of meander wavelength to suggesting that the lower Cayapas–Santiago estuary
channel width (L/W) allows characterization of a was also formed under tidal conditions. The last step
meandering river, the standard value for an equili- of a crevasse splay event is the formation of a new
brated stream in tropical environment ranging meandering pattern similar to the one observed
between 8 and 11 (Baker, 1978). The value of the upstream from the avulsion, but this step is not com-
(L/W) ratio is 15 for the channel segment upstream pleted in our case. The difference between our case
from point 1, and 13–17 between points 2 and 5 (not and a classical model of avulsion (Collinson, 1996;
considered the lowermost part influenced by tides). Smith et al., 1989) is that the evidence of changing
These homogeneous values suggest that similar con- fluvial pattern does not occur at the initiation point of
ditions of meander formation characterize the Cayapas the avulsion, but at the confluence between the San-
River upstream from point 1 and the Atajos River tiago and Cayapas rivers. This clearly means that the
along segments 2–5. These values are a few higher flood and subsequent avulsion have been generated by
than the standard values of Baker (1978). But accord- an excess discharge provided by the Santiago River.
ing to Dury (1970) this difference may be interpreted Usually a river avulsion ends by the partial or total
as a tendency to an underfit pattern, that is a wave- abandon of the previous river channel (Collinson,
length longer than the corresponding channel width, 1996; Smith et al., 1989), giving characteristic under-
due to a lower discharge than that of the equilibrated fit channel pattern if a reduced amount of water
pattern. However such a difference can also refer to supply is maintained in the old channel (Dumont,
local climate and sediment transport conditions 1996; Dury, 1970). In the present case the Los Atajos
(Schumm et al., 2000). These values are close to the River has not been abandoned and does not present an
9–15 ratio observed in the Amazon regions of north- underfit pattern. The interpretation emphasizes that
ern Peru (Dumont, 1991). These parameters show that the excess of water supply that generated the avulsion
the Atajos River represents the downstream continua- did not end as is the case for a flood, but has been
tion of the Cayapas River. This continuity of equili- definitive, because it is necessary to maintain enough
brated meandering pattern along the lower Cayapas water supply along the Los Atajos river to preserve
and the Los Atajos rivers suggests that the anasto- the former channel pattern.
mosed channel pattern observed downstream from the
confluence of the Cayapas and Santiago rivers is
triggered by an excess discharge or sediment transport 6. Morphologic pattern of the Cayapas–Santiago
issued from the Santiago River. delta
The morphology in point 2 suggests a channel
avulsion by a crevasse splay event initiated in the The Cayapas–Santiago and Mataje rivers join the
meander curve. Crevasse splay and avulsion are gen- sea through a coastal plain showing an ubiquitous
erated by flood, and the reconstruction of a new pattern of estuary and delta (Fig. 4). Previous studies
channel through the coastal plain involves successive qualified the Mataje River outlet of estuary, and the
steps analyzed by Smith et al. (1989) and synthesized Cayapas–Santiago of fluviodeltaic system (Tihay and
in Collinson (1996). At the first stage the splay Usselmann, 1995). The Mataje River estuary and the
expands rapidly into the marshes, forming a large Bolivar channel are located in the north part of the
avulsion belt with a high density of anastomosed delta; they are deep and accessible to sea ships. The
channels. This avulsion belt appears on the SPOT Bolivar Channel gives access to the San Lorenzo
image and aerial photos as an area of homogeneous harbor through a 5 to 18 m deep natural tidal channel
grain and colour (A and AV, Fig. 8). Then the network (INOCAR, 2002, unpublished data from San Lorenzo
of anastomosing channels gradually coalesce into harbor master’s office), except at the outlet to the sea
fewer and larger channels. This is the step observed where a sand ridge limits the low tide draw to 3 m (6–
on our example, and represented by the channel seg- 7 during upper tides). On the contrary the Cayapas–
ment 2–3 (Fig. 8). Downstream from point 3 the Santiago River outlet is almost closed at low tides by
114 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

a sub-emergent sand ridge, and presents in the outer


part the bottle pattern of a classical estuary (Perillo,
1995). According to Galloway (1975) and Haslett
(2000) these characteristics classify the area as a
compound wave and tide dominated delta. For
descriptive purpose the entire margin of beach ridges O: 5450-5130 BP
and mangrove crossed by several large tide channels
will be called Cayapas–Santiago deltaic coastal plain,
or simply coastal plain. The Cayapas–Santiago estu-
N: 5050-4860 BP
ary will refer only to the channel that crosses the outer
part of the coastal plain, as the inner part belongs to a 1000 m
fluvial system.
The coastal plain is 15 km wide and extends nearly
50 km north-eastward along the coast up to the
Colombian border. It includes two fluvial estuaries
at the outer parts (Cayapas–Santiago and Mataje riv-
ers) and two large tidal channels between them (Boca
de Limones and Bolivar channels) (Fig. 4). The tide
range in San Lorenzo is of about 3.5 m (mesotidal),
and the ebb–flood cycle is asymmetrical, with the ebb
8%–9% shorter than the flood (INOCAR, 2004),
favoring the extraction of suspended material. The
inner part of the Boca de Limones and Bolivar tidal
channels are sinuous and intricate, the main channels
ending sharply against the cliff bordering the coastal
plain. Small rivers connect the large tide channels to
the hinterland, the Bolivar channel ending in the
Estero El Salto, a typical dendrite shaped ria (Fig. 6).
The Mataje Fault bounds the coastal plain to the
north (Fig. 2), and a cliff to the north-east (Fig. 4, B1). Fig. 9. Extract of aerial photo localized on Fig. 4, showing partly
submerged beach ridges represented by the linear structures along
To the south-east a discontinuous scarp limits the
the central part of the photo. N and O number refer to calibrated
coastal plain from a 5–10 m high alluvial terrace of radiocarbon ages. See commentary in the text.
the Cayapas–Santiago River system (Winckell and
Zebrowski, 1997). To the south the coastal plain
narrows progressively ending at Las Peñas (Fig. 4). other evidences of the subsidence such as the deep
The delta morphology shows a striped pattern of tide channels observed to the north where the San
beach ridges which is very clear and continuous to the Lorenzo harbor is settled, and the typical ria pattern of
south but becomes discontinuous in the central part the Estero el Salto. The subsidence since about 5000–
and finally disappears to the North below mangrove 6000 BP is about equivalent to the tide range, that is
(Fig. 4). Partly drowned beach ridges are observed to 3.5 m (INOCAR, 2004), considering that the sea level
the W–SW of San Lorenzo (Fig. 9), showing coarse did not change significantly since that time. However,
beach sand and shell accumulation interpreted as the subsidence would be higher if the middle Holo-
upper tidal deposits that just outcrop during low cene sea level was higher than the present one, as
tide. A C14 dating on shell from the beach ridge suggested by Tihay (1989) and Tihay and Usselmann
facies gave a calibrated age of 5450–5130 BP. The (1995), and the sea level curve of Pirazzoli (1991) for
progressive drowning of the beach ridges from south- tropical areas. The clay, alluvium or volcanic con-
west to northeast gives evidence of a subsidence area glomerate of the Cachabi and San Tadeo formations
ending abruptly against the Mataje Fault. There are that underly the Holocene transgression of beach sand
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 115

is not a favorable material to support the interpretation points do not fit precisely with straight sections, prin-
of regional subsidence due to compacting, and the cipally in the northern part of the coastal plain, due to
increasing subsidence from south to north also favors field conditions (mangrove, marshes, tide channels).
the interpretation of a tectonic subsidence. However, the lateral continuity of the beach ridge
The hydrographic map drawn from SPOT image traces, especially in the south part of the coastal
and aerial photos shows that the coastal plain presents plain, allows interpolating dating using points that
two longitudinal parts, defined on the basis of tidal are not exactly on the section line, assuming that the
channel sinuosity and connections (dotted Z–ZVline, lateral extension of a beach ridge may be considered
Fig. 4). The inner part presents the higher density of approximately as a time mark. We used 11 dates of
tidal channels, and the outer part a more homogeneous beach ridges from Tihay and Usselmann (1995),
and relatively simple pattern, with all the small tidal sampled in the southern part of the littoral zone
channels originating near the inner limit of the outer (Fig. 4, points A to G and I to L). These dates were
part. Crossing the Z–ZVline the channel pattern of the made on shells from the ridges, recovered from 1 to 2
Santiago–Cayapas estuary changes significantly. m deep holes. The 8 other dates (Fig. 6, points H and
Through the inner belt the channel is relatively narrow M to Q) come from this study. Seven dates are
with a poorly anastomosed pattern, but it suddenly obtained on organic material (small wood fragments
widens, showing an inflexion or the trend to the west or leaves accumulation) recovered at a depth of 1–2 m
in the outer part. Also the Los Atajos River, old by vibro-coring in the wetlands of ridge slack depos-
segment of the Cayapas river, reaches the Boca de its, and one (sample O) comes from shell recovered in
Limones tidal channel near the Z–ZVlimit (Fig. 1) in beach ridge deposit recovered during low tide.
an intricate area of broadening and merging of differ- We checked the homogeneity of dating for the two
ent channels. A careful observation of the aerial photo types of material by comparing the age of sample O
and the SPOT image shows that the beach ridges get on shell from a beach ridge deposit with sample N
straight on the channel border in the inner part of the from organic matter in slack deposit close to the ridge
deltaic coastal plain, but present a turn toward the (Fig. 9). The ages are 5450–5130 BP for sample O
upstream direction in the outer part (Fig. 8). This and 5050–4860 BP for sample N, that is 300–400
observation suggests that the inner part of the channel years less for the sample located 200 m on the seaside.
cuts through pre-existing beach ridges, when the outer This result is coherent considering the 1–2 m y- 1
wide channel was formed as a classical seaward mean rate of beach ridge accretion calculated for
accretion delta. In other words, the present inner and this area (see below). The oldest age (6880–6660
narrow channel of the Santiago–Cayapas estuary has BP, P) comes from a tree trunk found in the innermost
been opened through a pre-existing belt of beach part of the Boca de Limones channel, near the foot of
ridges by the time the shore line was near the Z–ZV the littoral cliff. This trunk is only partly emerged
limit (Fig. 4). The later evolution of the coastal margin during very low tide. Along the present beaches the
built a classical bottle shaped estuary (Perillo, 1995). tree trunks that are brought to the shore are preserved
only in the upper part of the beach, near the top of the
beach ridges, or on their backside. The present posi-
7. The pattern of beach ridge accretion tion of the old tree trunk imply a subsidence of at least
the value of the tide range (3.5 m) since 6880–6660
The beach ridges are considered as an opportunity BP. The radiocarbon ages of the inner part of the
to date the coastline when the avulsion event occurred coastal plain range between 6880–6660 BP (P) and
that formed the new Cayapas–Santiago channel (Fig. 4132–3686 BP (K), and those of the outer part
2, point 3). We also analyzed the succession of the between 460–0 BP (H) and 3260–2830 BP (E). The
beach ridges in order to get possible evidence of D–E and M samples which are laterally distant of
irregularities in this process of accretion in relation about 20 km along the same stripe of ridges have
with the avulsion event. similar ages. The dates have been plotted as a function
The succession of beach ridges is determined using of their distance from the shoreline (Fig. 10A) or from
19 radiocarbon dating, located on Fig. 10. The dating the inner cliff (Fig. 10B). Two diagrams are used to
116 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

7000
X3
Slow P
accretion
6000
J I
X1
5000
X4 O
Age in years

Fast N
accretion
4000 L K
E FA X2
D
3000
Slow M
accretion G Y1
2000
F
1000 A
B
H
Y2 B
0
0 5 10 15 km
Distance from the shoreline

Slow
accretion X3 7000
400 yr margin of error P
X1 6000
J I
Fast O 5000
accretion N X4

Age in years
Y1 FA L
K
4000

X2
D M 3000
G
Slow
2000
accretion
F
Shoreline

Y2 1000

A 0
15 km 10 5 0

Distance from the cliff

Fig. 10. Diagrams of the beach ridge construction of the littoral margin, using age and positions from Fig. 4, with the same letters. Isolated data
such as Q, and dating problem such as C are not used. The dotted line represents the mean accretion of the littoral margin. The lines X1–X2, X2–
X3 and Y1–Y2 associate data along sections positioned on Fig. 4, and represented respectively by open (X1–X2), grey (X3–X4, and black bar
lines ( Y1–Y2). The length of bar lines represents the margin of error. A: age versus distance measured from the present shoreline, and using all
the data. B: age versus distance measured from the inner cliff. The black point represents the present position of the shoreline along the Y1–Y2
segment.

take into account the fact that the shoreline is not between them. Sampling position allows definition
exactly parallel to the inner cliff, the uncertainty of two sections through the inner margin (Fig. 10:
increasing with the distance; however the two dia- X1–X2 and X3–X4) and one through the outer one
grams present very similar patterns. The diagrams (Fig. 10: Y1–Y2). Across the inner margin the X1–X2
give evidence of a mean rate of beach ridge accretion section shows mean rates of beach ridge construction
of 2–2.5 m y 1 since about 6000 BP (dotted line Fig. of 1.75 m yr 1 between 5935–5581 BP (I) and 4230–
10). However a more detailed analysis suggests 3 3806 BP (L), and the section X3–X4 of 0.7–1 m yr 1
groups of points, defining two periods of relatively between: 6880–6660 BP (P) and 5050–4860 BP (N).
slow accretion and one period of faster accretion Data from the outer margin along section Y1–Y2 show
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 117

first a faster construction at a rate of up to 5 m yr 1


between 3251–2854 BP (D) (or 3200–2780 BP, M)
and 2781–2340 BP (G), decreasing to about 1 m yr 1 N
during the more recent period. There is no data in the
time range 4200–3200 BP (data L and M, D or E,
Figs. 4 and 10), the pattern of the diagram suggesting
a discontinuity more than a progressive change.

8. Evidence of Holocene tectonics: the San Lorenzo


Fault

South-west and north of San Lorenzo a sharp


lineament observed on the SPOT image and aerial
photos defines the limit between the coastal plain
and the San Lorenzo upland (Figs. 4, 11, 12). On
the field this corresponds to a 10–20 m high scarp
visible over more than 10 km. This scarp has been 1000 m
previously identified and interpreted as an active fault
(Eguez et al., 2003; Santana and Dumont, 2002), for
the sharp limit with the littoral plain, but precise
description and argument for its active motion have A
never been presented.
The scarp line cuts the reddish weathered detritus
of the Cachabi Formation. North of San Lorenzo the
lineament splits into two branches (Fig. 4). The main P: 6880-6660 BP
scarp follows the eastern branch, but disappears pro-
gressively in the upland. North of San Lorenzo the
western segment is a tiny line across mangrove and
wetland (Fig. 12), but becomes sharper to the north,
making again the border of the upland when the
eastern line disappears (Fig. 4).
The usual question with such coinciding fault and
Fig. 11. Extract of aerial photo localised on Fig. 4, showing the San
coast line is whether the scarp line represents a fault
Lorenzo Fault (A) between the mangrove and tidal channel to the
line that post-dates and cuts the mid-Holocene trans- west and the upland to the east. P refers to the radiocarbon age of a
gression, or a beach angle that uses locally a former tree trunk.
fault scarp. We observe here some ubiquitous ele-
ments, suggesting that different events–transgression
and fault–may have occurred more or less simulta- Faults through the coastal plain of mangrove (Fig. 12)
neously. For example the conformity between the is considered as a determining element, to interpret
trend of the inner beach ridges and the scarp suggests this lineament as a fault motion that necessarily post-
a shore angle of the postglacial sea rise, but also the dates the mid-Holocene transgression. We did not find
line is too continuous and sharp on aerial photos to be evidence of beach deposits on the upland side of the
just a 6000-year-old fossil sea cliff (Fig. 11). In parti- San Lorenzo Faults to attest the post-transgression
cular, no weakening or deviation of the line appears motion of the fault scarp. The interpretation suggested
near the lows and small valleys across the cliff, as here is that the 20-m scarp and lineament of the San
observed when a shoreline meets estuaries. Finally the Lorenzo Fault probably predates the post-glacial
trace of the north-western segment of the San Lorenzo transgression, but has been reactivated later.
118 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

Fault
N not synchronous, but they belong to the same family
and are representative of the more recent and probably
active fault activity of the area.

9. Discussion

9.1. A succession of related morphologic events

Three important morphological changes have been


identified: 1–the diversion of the Mira River to the
north at the head of the El Placer fan, 2–the diversion
of the drainage network to the south-west in the lower
part of the fan, and 3–the channel pattern change
along the Cayapas–Santiago river segment. The ana-
lysis of beach ridges accretion in the coastal plain is a
complement of point 3. It is possible to consider all
1000 m these morphological changes as independent. How-
ever, there is some coherency and relation from one
Fig. 12. Extract of aerial photo localized on Fig. 4, showing the San change to the other, and the hypothesis supported here
Lorenzo Fault through a wetland of mangrove and tidal channels.
is a link between them. The diversion of the Mira
River in Alto Tambo, at the head of the fan, presents
The fault plane of the eastern segment of the San the same style of diversion than the one observed in
Lorenzo Fault has been observed near San Lorenzo the lower part of the fan in La Boca. It is considered as
(point B1, Fig. 4A). It trends N40–608E and dips of an early event inside the same structural scheme of
708 to 808 to the north-west, and cuts the brown- evolution, but not synchronous of the events that
reddish sandy clay of the Cachabi Formation (Fig. 4 occurred at the lower part of the fan.
B1). Associated fault planes from the same area trend The drainage diversion in La Boca implies the
E–W (Fig. 4 B1, fault planes numbers 5 and 7) and presence and the motion of the La Boca Fault,
dip generally to the north, with some associated which is sub-parallel to the San Lorenzo Faults. The
planes dipping to the south. The calculation of the La Boca and the San Lorenzo faults border the La
main stress axis with the Carey method (Carey, 1979; Boca uplift, the regional structure responsible for the
Carey and Mercier, 1987) gives a maximum stress re-routing of the lower part of the El Placer fan
direction r1 nearly vertical, and r2 and r3 nearly drainage toward the south-west. The north-eastern
horizontal trending N808E and N1708E, respectively. segment of the San Lorenzo Faults presents a trans-
The slickensides (Fig. 4 B1) present a relative disper- tension motion (dextral-normal), and a similar motion
sion that is explained by the high dip of the fault can be easily hypothesized for the La Boca Fault. The
planes. The statistical analysis shows that the angle diversion of the Palabi, Tululbi, Bogotá and Carolina
between the measured and calculated slickensides is rivers to the Southwest against the La Boca uplift, and
lower than 58 for most of the fault planes, and less their re-routing toward the Santiago River, should
than 118 for all, and thus the result may be considered have increased the discharge of this river. However
as coherent. The conclusion of the Holocene tectonic this increase of the discharge is accompanied by a
activity is a nearly N–S extension with a transtension decrease of the slope along the diverted river seg-
motion along E–W to NE–SW trending faults. The ment, meaning that the discharge increase cannot be
San Lorenzo faults trend parallel to the faults correlated with an increase of the sedimentary charge
observed on the south-eastern border of the La Boca of the Santiago River. This excess discharge is con-
uplift, and also parallel to the North Canandé fault in sidered as responsible for the channel pattern change,
Alto Tambo. The motion of all these faults is probably from meandering to straight, observed downstream
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 119

the confluence between the Santiago and Cayapas the Cayapas–Santiago estuary changes from a fluvial
rivers, and the avulsion that created downstream the to an estuary pattern (Fig. 8, point 3) is indicative of
new Cayapas–Santiago estuary through the coastal the age of the drainage diversion in La Boca and the
plain. The dual pattern of the Cayapas–Santiago estu- resulting avulsion of the Cayapas–Santiago River.
ary through the coastal plain supports this interpreta- This age can be determined using the striped pattern
tion: the narrow upstream part has the pattern of a of beach ridges (Figs. 8 and 4). The D sample (3251–
river avulsion channel opened through the pre-exist- 2854 BP) (Fig. 4) is the closest point, the sudden
ing coastal margin of beach ridges, and the wide widening of the channel beginning about 200 m sea-
downstream part has the classical tide-wave domi- ward. Also the extrapolation of beach ridge traces on
nated pattern of estuary. Fig. 5 crosses the point of channel change at about
We consider that the motion of the La Boca fault 3000–3200 BP.
created the favorable situation for the increase of the
discharge, but the precise time correlation between 9.3. Post-seismic deformation of the coastal plain?
fault motion and the onset of the avulsion event may
be discussed. An avulsion characterizes a peak of Before 4000 BP the rate of beach ridge construc-
discharge, that is also generally a peak of precipita- tion is slow (0.7–1.7m/yr). After that we lack data in
tion elsewhere in the catchment area. In this area the the time range 4200–3200 BP. A period of fast
highest precipitation occurs generally during El Nino accretion (5 m/yr) appears after 4000–3600 BP, and
periods, with two to three times more rain than during in the time range 3200–2700/2700–2300 BP (Fig.
normal periods, and concentration of the rain during 10). The transition from the early period of slow
short periods (Perrin et al., 1998). That means that the accretion to the period of fast accretion is not docu-
avulsion may have occurred during a period of high mented, but no progressive change is perceptible.
precipitation following the re-routing of the drainage From 2000 BP to present we observe a progressive
along the La Boca uplift. However special attention to slow down of the accretion rate (about 1 m/yr), but
the pattern of the avulsion leads us to discard a simple the present rate (during the last century for example)
climate induced avulsion: after the avulsion there is is not documented.
no abandonment or even notable reduction of the The faster beach ridge accretion may be due
previous channel (the Los Atajos River). This either to an excess of sediment supply to the coast,
means that the discharge increase that generated the or to vertical motion of the coastal plain due to co-
avulsion did not really disappear after that event. A or post-seismic deformation. We have observed that
peak of precipitation and flood may have triggered the river diversion against the La Boca uplift corre-
the avulsion, but the return to normal conditions sponds to a lower slope than the previous one, and
would have left an underfit pattern for the Los Atajos thus cannot be correlated with an increase of the
river, as observed for the avulsion events in the sedimentary supply to the coast. Short-term climatic
western Amazonian basin (Dumont, 1996). This variations may be related to variation of the beach
means that the discharge increase is definitive, and ridge construction but at a higher frequency of var-
that a climate induced flood event is not the explana- iation than the progressive change observed over a
tion. This interpretation is coherent with the fact that 3000-year period. Our data suggests to correlate the
the channel pattern change which is associated with fast beach ridge accretion in the time range 3200–
the avulsion began at the confluence between the 2300 BP with a co- or post-seismic uplift of the
Santiago and Cayapas rivers, and not just at the coastal plain. This is more probably a post-seismic
avulsion point as is usually the case (Collinson, motion because of the duration of the event and the
1996; Smith et al., 1989). progressive variation to the recent period of slow
accretion. In a low sloping shore platform a small
9.2. Age of the drainage disruption uplift can give account of the formation of beach
ridges some hundreds of meters seaward. The pre-
According to the proposed scenario the age of the sence of emerged banks of dead oysters in life
beach ridges at the point where the channel pattern of position (Fig. 4) supports this hypothesis.
120 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

9.4. Relation with the seismic zone rial. However drainage network is generally dense and
active, and may be a suitable morphological tool. New
The seismic activity of the Esmeraldas–Tumaco faults showing recent or active motion have been
area is originated in the subduction zone (Collot et identified on the basis of morphologic analysis, the
al., 2004, 2002; Herd et al., 1981). The motion of the North Canandé, La Boca and Mataje faults, and pre-
Las Boca and San Lorenzo faults cannot be directly cision has been brought on the Holocene motion of
linked to the seismogenic zone, but only to the accom- the San Lorenzo fault. In these cases the morphologic
modation of the deformation in the upper crust of the expression of active tectonics seems enhanced by the
overriding plate. The comparison of stress tensor in change from shortening to extension tectonics in the
the subduction zone (Ego et al., 1996) and in the coastal margin that occurred during the Quaternary,
upper plate in the San Lorenzo fault (Fig. 4) supports and particularly the recent N–S extension, that is
this interpretation. The state of stress obtained from clearly associated with the disruption of the lower
the inversion of focal solutions from subduction earth- part of the Pleistocene El Placer fan.
quakes (Fig. 4 B2) shows a ENE–WSW shortening The regional evolution of the drainage network in
(Ego et al., 1996). The comparison with the state of the San Lorenzo area describes the disruption of the
stress of the San Lorenzo Fault (Fig. 4 B1) suggests a drainage network of the Pleistocene El Placer fan,
permutation according to r1Yr2; r2Yr3 and diverted to the north at the head of the fan and to
r3Yr1, that is coherent with an accommodation of the southwest in the central part. This southwestward
the deformation in the upper plate, favored by the diversion increases downstream the discharge of the
oblique subduction. This structural relation allows Santiago River, and triggered the avulsion of the
discarding the effect of gravitational accommodation Cayapas–Santiago River through a new estuary
in the upper inner edge of the continental shelf, opened across the coastal plain of beach ridges. The
because the motion should be represented by E–W avulsion is dated 3200–2800 BP by the beach ridges
or NW–SE trending extension instead of the observed located at the point of changing pattern of the estuary,
north–south one. showing anastomosed pattern for the avulsion channel
The seismic activity along the South American in the inner part and the new tide-wave dominated
Pacific margin is frequently accompanied by co-seis- estuary in the outer part. The avulsion event will be
mic deformation (Barrientos, 1996; Barrientos and called bLa Tolita eventQ from the position of the La
Plafker, 1992). In the Esmeraldas-Tumaco seismic Tolita archeological site at the point where the avul-
zone similar vertical motion have been mentioned, sion channel reaches the coast.
but have never been precisely quantified and located The disruption of the El Placer fan and the related
relative to the earthquakes. The available information morphological changes are clearly related to fault
from the Tumaco area reports generally co-seismic motion. The diversion to the north of the Mira
subsidence (Herd et al., 1981), but also uplift adjust- River at the head of the fan is related to the motion
ment that seems to occur after the earthquake, and of the North Canandé Fault and other N–S trending
reaching 33 mm in 28 years (Tihay and Usselmann, faults. The motion of the La Boca Fault diverted the
1998). This data supports the hypothesis of a post drainage of the central part of the fan toward the
seismic uplift of the shore margin in the central part of Santiago River, evidencing the La Boca uplift. This
the Cayapas–Santiago delta. uplift is bordered to the NW by the San Lorenzo
Fault, that makes the border of the coastal plain and
presents evidences of motion since the mid-Holo-
10. Conclusion cene. These morphologic-fault relationships are inter-
preted as the effect of one or several strong
The San Lorenzo area represents a good example earthquakes in the time range 3200–2800 BP. How-
of how to provide evidences of active tectonics under ever the northward diversion of the Mira River pre-
wet tropical regions. The tropical climate is not sui- dates the river diversion observed against the La
table for the preservation and the observation of fault Boca Fault, suggesting that the tectonic changes
scarps, especially with low topography and soft mate- that disrupted the El Placer fan migrated from the
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 121

upper part of the western Andean Cordillera to the realized in cooperation with the bVariabilidad CosteraQ
foothills and the littoral margin. INOCAR (Instituto Oceanográfico de la Armada,
Additional and independent elements come from Ecuador) project. We are grateful to Edgar Rivas
the analysis of the pattern of beach ridges accretion. A (INOCAR) and Kervin Chunga (University of Guaya-
period of fast beach ridges accretion apparently quil) for their help during field work, and to the
accompanied and continued after the 3200–2800 La personal of the San Lorenzo Harbor Master’s Office
Tolita event. The transition to the recent period of for their help in logistic organization of field work.
slow accretion is progressive. This progressive change The final manuscript has benefited from comments
in beach ridge accretion cannot be related to climate and suggestions from Martin Stokes and Edward A.
oscillations or increase of the sediment supply after Keller, for which they are thanked. We thank Ben
the avulsion event, but to small uplift of the coastal Yates for the final correction of the text. This study
margin in relation to one event or a crisis of seismic participates in the activity of the IGCP project 495,
events. The present period documented for the last Land Ocean Interaction. This is publication 756 of
century is characterised by a high seismic activity, but UMR 6526-Geosciences Azur.
we do not know precisely what is the present rate of
beach ridge accretion in the area. This precision will
help to specify whether the present historic period that References
includes the 1906 Tumaco Mw 8.8 earthquake may be
compared or not with the previous 3200–2800 BP Aalto, K.R., Miller, W., 1999. Sedimentology of the Pliocene Upper
period of fast ridge accretion. Onzole Formation, an inner-trench slope succession in north-
western Ecuador. Journal of South American Earth Sciences 12
Finally a tentative relation can be made between the (1), 69 – 85.
morphologic and cultural evolution of the area. It is Alvarado, A., 1998. Variation du champ de contrainte et de défor-
worth noting the particular position of the La Tolita site, mation et quantification des déformations actives du bloc côtier
at the point of the coast where the avulsion channel has de l’Equateur. DEA de Géodynamique et physique de la terre
been opened (Fig. 4). The pre-Colombian occupation Thesis, Paris XI, centre d’Orsay, Orsay, 54 pp.
Baker, V.R., 1978. Adjustment of fluvial system to climate and
of the studied area is dominated by the La Tolita source terrain in tropical and subtropical environments. In:
culture, one of the most important coastal occupations Miall, D. (Ed.), Fluvial Sedimentology. Canadian Society of
along the South American coast (Bouchard, 1995; Petroleum Geology, pp. 211 – 230.
Valdez, 1987). The settlement in La Tolita began Baldock, J.W., 1982. Geologı́a del Ecuador: Boletı́n de la explica-
about 3000 BP, and the peak of the development ción del Mapa Geológico de la República del Ecuador, Esc.
1 : 1V000.000. Min. Rec. Nat. Energ., Quito, 10., Quito.
occurred between 2250 and 2000 BP (Valdez, 1987). Barrientos, S.E., 1996. On Predicting Coastal Uplift and Subsidence
According to Valdez (1987) the location of the La Due to Large Earthquakes in Chile, Third ISAG. ORSTOM, St
Tolita site in the Cayapas–Santiago estuary, close to Malo, France, pp. 145 – 148.
the Pacific Ocean, is the result of a specific strategy to Barrientos, S.E., Plafker, G., 1992. Postseismic coastal uplift in
get easier access to areas along the littoral, but also southern Chili. Geophysical Research Letters 19, 701 – 704.
Beck, S.L., Barrientos, S., Kausel, E., Reyes, M., 1998. Source
towards the hinterland. In the case of La Tolita the new characteristics of historic earthquakes along the central Chile
drainage network of the Santiago River may have made subduction zone. Journal of South American Earth Science 11
the gold bearing area more accessible, thus increasing (2), 115 – 129.
the capacity of acquiring one of its major resources. Bès de Berc, S., Soula, J.C., Baby, P., Souris, M., Christophoul, F.,
Rosero, J., 2005. Geomorphic evidence of active deformation
The seaport that was also now available afforded the
and uplift in a modern continental wedge-top-foredeep transi-
coastline trade along the Pacific and gave the La Tolita tion: example of the eastern Ecuadorian Andes. Tectonophysics
center its cultural reputation and economic hegemony. 399, 351 – 380.
Bishop, P., 1995. Drainage rearrangement by river capture,
beheading and diversion. Progress in Physical Geography 19
Acknowledgements (4), 449 – 473.
Booth-Rea, G., J.-M., A., Azor, A., Garcia-Duenas, V., 2004.
Influence of strike slip fault segmentation on drainage evolution
This study is part of the DEMA3-IRD (Institut de and topography. A case study: the Palomares Fault Zone (south-
Recherche pour le Développement, France) project, eastern Betics, Spain). Journal of Structural Geology, in press.
122 J.F. Dumont et al. / Geomorphology 74 (2006) 100–123

Bouchard, J.F., 1995. Altas Culturas y Medio Ambiente en el Dumont, J.F., Lavenu, A., Ortlieb, L., Guillier, B., Alvarado, A.,
Litoral Norte del Area Ecuatorial Andina, Cultura y Medio Benitez, S., Jouannic, C., Martinez, C., Labrousse, B., Poli, J.T.,
Ambiente en el Area Andina Septentrional. Abya-Yala, Quito, 1997. Extensional tectonics in the coastal block of Ecuador:
pp. 195 – 223. preliminary results and implications. Workshop on Late Qua-
Bowler, J.M., Harford, L.B., 1966. Quaternary tectonics and the ternary Coastal Tectonics, London.
evolution of the Riverine Plain near Echuca, Victoria. Geologi- Dumont, J.F., Santana, E., Vilema, W., 2005. Morphologic evidence
cal Society of Australia 13, 339 – 354. of active motion of the Zambapala Fault, Gulf of Guayaquil
Carey, E., 1979. Recherche des directions principales de contra- (Ecuador). Geomorphology 65, 223 – 239.
intes associées au jeu d’une population de failles. Revue de Dury, G.H., 1970. General theory of meandering valleys and under-
Géographie Physique et de Géologie Dynamique 21 (1), fit streams. In: Dury, G.H. (Ed.), River and River Terraces. Mc
57 – 66. Millan, London, pp. 264 – 275.
Carey, E., Mercier, J.L., 1987. A numerical model for determining Ego, F., Sebrier, M., Lavenu, A., Yepes, H., Eguez, A., 1996.
the state of stress using focal mechanisms of earthquake Quaternary state of stress in the Northern Andes and the
population: application to Tibetan teleseims and microseismi- restraining bend model for the Ecuadorian Andes. Tectonophy-
city of Southern Peru. Earth and Planetary Science Letters 82, sics 259, 101 – 116.
165 – 177. Eguez, A., Alvarado, A., Yepez, H., Machette, M.N., Costa, C., Dart,
Carver, G.A., McCalpin, J., 1996. Paleoseismology of compres- R.L., 2003. Database and map Map of Quaternary faults and folds
sional tectonic environments. In: McCalpin, J. (Ed.), Paleoseis- of the Ecuador and its offshore regions. International Lithosphere
mology. International Geophysics Series. Academic Press, Program Task Group II-2, Major Active faults of the world, U. S.
London, pp. 183 – 270. Department of the Interior, U. S. Geological Survey, Denver.
CERESIS, 1985. Mapa Neotectonico Preliminar de America del Evans, C.D.R., Whittaker, J.E., 1982. The geology of the western
Sur. part of the Borbón Basin, North west Ecuador: Trench forearc
CODIGEM, 1993. Mapa Geológico de la República del Ecuador, Geology. Geological Society of London 10, 191 – 200.
escala 1/1 000 0008. British Geological Survey. Fitch, T.J., Scholz, C.H., 1971. Mechanism of underthrusting in
Collinson, J.D., 1996. Alluvial sediments. In: Reading, H.G. (Ed.), Southwest Japan: a model of convergent plate interactions.
Sedimentary Environments: Processes, Facies and Stratigraphy. Journal of Geophysical Research 76 (29), 7260 – 7292.
Blackwell Science, London, pp. 37 – 82. Galloway, W.E., 1975. Process framework for describing the mor-
Collot, J.-Y., Charvis, P., Gutscher, M.-A., Operto, S., 2002. Explor- phologic and stratigraphic evolution of deltaic depositional
ing the Ecuador–Colombia active margin and interplate seismo- systems. In: Broussard, M.L. (Ed.), Deltas: models for Explora-
genic zone. EOS, Transactions, American Geophysical Union, tion. Huston Geological Society, Huston, pp. 87 – 98.
83(17): 185, 189–190. González, J.L., Correa, I.D., Aristizábal, O., 2002. Evidencias de
Collot, J.Y., Marcaillou, B., Sage, F., Michaud, F., Agudelo, W., subsidencia cosı́smica en el delta del San Juan. In: Correa,
Charvis, P., Graindorge, D., Gutscher, M.-A., Spence, G., 2004. I.D.a.R.J.D. (Ed.), Geologı́a y Oceanografı́a del delta del Rı́o
Are rupture zone limits of great subduction earthquakes con- San Juan, Litoral Pacifico Colombiano. Universidad EAFIT,
trolled by upper plate structures ? Evidence from multichannel Medellı́n, pp. 91 – 110.
seismic reflection data acquired across the Northern Ecuador– Gutscher, M.A., Malavieille, J.S.L., Collot, J.-Y., 1999. Tectonic
southwest Colombian margin. Journal of Geophysical Research segmentation of the North Andean margin: impact of the Car-
109, B11103. negie Ridge collision. Earth and Planetary Science Letters 168,
De Mets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current 255 – 270.
plate motions. Geophys. J. Int. 101, 425 – 478. Haslett, S.K., 2000. Coastal Systems. Routledge Introduction to
Deniaud, Y., 2000. Enregistrements sédimentaire et structural de Environment. Routledge, London. 218 pp.
l’évolution géodynamique des Andes Equatoriennes au cours du Herd, D.G., Youd, T.L., Meyer, H., Arango, C.J.L., Person, W.J.,
Néogène: Etude des bassins d’avant arc et bilan de masse. Mendoza, C., 1981. The Great Tumaco Colombia earthquake of
Géologie Alpine, Mémoire HS (32), 159. 12 December 1979. Science 211, 441 – 445.
Dumont, J.F., 1991. Fluvial shifting in the Ucamara Depression, as Huang, W., 1993. Morphologic patterns of stream channels on the
related to neotectonics of the Andean foreland–Brazilian craton active Yishi Fault, southern Shandong Province, Eastern China:
border. Géodynamique 6 (1), 9 – 20. implication for repeated great earthquake during the Holocene.
Dumont, J.F., 1993. Lake pattern as related to neotectonics in Tectonophysics 219, 283 – 304.
subsiding basins: the example of the Ucamara Depression, IGM, 1991. Esmeraldas, Hoja NII-NA 17-15. Instituto Geográfico
Peru. Tectonophysics 222, 69 – 78. Militar, Quito.
Dumont, J.F., 1996. Neotectonic of Subandes–Brazilian craton IGM, 1992a. Ibarra, Hoja ÑII,NA, 17,18,16. Instituto Geográfico
boundaries using geomorphological data: the Marañon and Militar, Quito.
Beni Basins. Tectonophysics 257, 137 – 151. IGM, 1992b. San Lorenzo, Hoja ÑI, NA 17-12. Instituto Geográfico
Dumont, J.F., Fournier, M., 1994. Geodynamic environment of Militar, Quito.
Quaternary morphostructures of the subandean foreland basins INOCAR, 2004. Tablas de mareas y datos astrónomicos del sol y la
of Peru and Bolivia: characteristics and study methods. Qua- luna. República del Ecuador, Instituto Oceanográfico de la
ternary International 21, 129 – 142. Armada, Guayaquil. 117 pp.
J.F. Dumont et al. / Geomorphology 74 (2006) 100–123 123

Jackson, J., Norris, R., Youngson, J., 1996. The structural evolution PRONAREG-ORSTOM, 1984. Valdez, Mapa Morfo-Pedológico.
of active fault and fold systems in central Otago, New Zealand: Ministerio de Agricultura y Ganaderı́a, Quito.
evidences revealed by drainage patterns. Journal of Structural Santana, E., Dumont, J.F., 2002. The San Lorenzo Fault, a new
Geology 18, 217 – 234. active fault in relation to the Esmeraldas–Tumaco seismic zone.
Kelleher, J.A., 1972. Ruptures zones of large South American 5th International Symposium on Andean Geodynamics. IRD,
earthquakes and some predictions. Journal of Geophysical Toulouse, pp. 577 – 580.
Research 77 (11), 2087 – 2103. Santana, E., Dumont, J.F., King, A., 2001. Los efectos del feno-
Keller, E.A., Pinter, N., 2002. Active tectonics: Earthquakes, Uplift meno El Niño en la ocurencia de una alta tasa de erosion costera
and Landscape. Prentice Hall, Upper Saddle River. 206 pp. en el sector de Punta Gorda, Esmeraldas. Acta Oceanografica
Kellogg, J.N., Vega, V., 1995. Tectonic development of Panama, del Pacifico 11 (1), 1 – 5.
Costa Rica, and the Colombian Andes: constraints from Global Scheu, E., 1911. Le grand tremblement de terre de la Colombie
Positioning System geodetic studies and gravity. In: Mann, P. (Monographie de quelques grands séismes de l’année 1906).
(Ed.), Geologic and Tectonic Development of the Caribbean Catalogue Régional des Tremblements de Terre Ressentis Pen-
Plate Boundary in Southern Central America, Geol. Soc. Am. dant L’année 1906, Strasbourg, France, pp. 36 – 44.
Spec. Pap., vol. 295, pp. 75 – 90. Schumm, S.A., Dumont, J.F., Holbrook, J.M., 2000. Active Tec-
McCalpin, J. (Ed.), Paleoseismology. International Geophysics Ser- tonics and Alluvial Rivers. Cambridge University Press, Cam-
ies, vol. 62. Academic Press, London. 588 pp. bridge. 276 pp.
Mörner, N.A., 2003. Paleoseismicity of Sweden, a Novel Paradigm. Smith, N.D., Cross, T.A., Dufficy, J.P., Clough, S.R., 1989. Anat-
Jofo Grafiska AB, Stockholm. 320 pp. omy of an avulsion. Sedimentology 36, 1 – 23.
Ortlieb, L., Barrientos, S., Guzman, N., 1996. Coseismic coastal Swenson, J.L., Beck, S.L., 1996. Historical 1942 Ecuador and 1942
uplift and coraline algae record in northern Chile: the 1995 Peru subduction earthquakes, and earthquakes cycles along
Antofogasta earthquake case. Quaternary Science Reviews 15, Colombian–Ecuador and Peru subduction segments. PAGEOPH
949 – 960. 146 (1), 67 – 101.
Pedoja, K., 2003. Les terrasses marines de la marge Nord Andine Tihay, J.P., 1989. Aspects Geomorphologiques de L’environnement
(Equateur et Nord Pérou): relations avec le contexte géodyna- du site Archéologique de la Tolita (Equateur). Universite de Pau
mique. Ph.D Thesis, Université Pierre et Marie Curie, Paris, et des Pays de l’Adour, Pau.
350 pp. Tihay, J.P., Usselmann, P., 1995. Medio ambiente y ocupación
Pedoja, K., Dumont, J.F., Sorel, D., Ortlieb, L., 2001. Marine humana en el litoral Pacı́fico Colombo-ecuatoriano. In: Guinea,
Terraces and Subducting Asperities: the Manta Case, Ecuador, J.F.B.y.J.M.M. (Ed.), Cultura y Medio Ambiente en el Area
Fifth International Conference on Geomorphology. Transaction Andina Septentrional. Abya-Yala, Quito, pp. 377 – 399.
of the Japanese Geomorphological Union, Tokyo, p. 187. Tihay, J.P., Usselmann, P., 1998. Ambientes humedos de la costa
Perillo, G.M.E. (Ed.), Geomorphology and Sedimentology of Estu- pacifica ecuatorial (Colombia y Ecuador) y uso antropico; geo-
aries. Development in Geomorphology, vol. 53. Eslsevier. dinamica y aportes de los sensores remotos. In: Mercedes
471 pp. Guinea, J.M.a.J.F.B. (Ed.), El Area Septentrional Andina.
Perrin, J.L., Jeanneau, J.L., Podwojeski, P., 1998. Deslizamientos de Abya-Yala, Quito, pp. 67 – 80.
Tierra, Inundaciones y Flujos de Lodo en Esmeraldas; Diagnos- Valdez, F., 1987. Proyecto Archeologico bLa TolitaQ. Museos del
tico General de la Situacion Actual en la Ciudad, Mision de Banco Central del Ecuador, Quito. 91 pp.
Expertos. Informe ORSTOM, Quito. 020005212. Winckell, A., Zebrowski, C., 1997. Los paisajes costeros. In:
Pirazzoli, P.A. (Ed.), World Atlas of Holocene Sea-Level, vol. 58. Winckel, A. (Ed.), Los Paisajes Naturales del Ecuador. Geogra-
Elsevier Oceanography Series, Amsterdam. 171 pp. fı́a Básica del Ecuador. CEDIG, Quito, pp. 208 – 319.
Potter, P.E., 1978. Significance and origin of big rivers. Journal of Witt, C., 2001. Análisis de la deformación reciente y potencialmente
Geology 86, 13 – 33. activa con base a imágenes radar, fotos aéreas, DEM y obser-
PRONAREG-ORSTOM, 1984a. Ibarra, Mapa Morfo-Pedológico. vaciones microtectónicas en la provincia de Esmeraldas, Ecua-
Ministerio de Agricultura y Ganaderı́a, Quitp. dor. Tesis de ingeniero Thesis, Escuela Politécnica Nacional,
PRONAREG-ORSTOM, 1984b. Tulcan, Mapa Morfo Pedológico. Quito, 122 pp.
Ministerio de Agricultura y Ganaderı́a, Quito.

You might also like