Velandia2018 PDF

You might also like

You are on page 1of 58

Accepted Manuscript

The transpressive southern termination of the Bucaramanga fault (Colombia): Insights


from geological mapping, stress tensors, and fractal analysis

Francisco Velandia, Mauricio A. Bermúdez

PII: S0191-8141(18)30095-6
DOI: 10.1016/j.jsg.2018.07.020
Reference: SG 3715

To appear in: Journal of Structural Geology

Received Date: 15 February 2018


Revised Date: 21 July 2018
Accepted Date: 21 July 2018

Please cite this article as: Velandia, F., Bermúdez, M.A., The transpressive southern termination of the
Bucaramanga fault (Colombia): Insights from geological mapping, stress tensors, and fractal analysis,
Journal of Structural Geology (2018), doi: 10.1016/j.jsg.2018.07.020.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
1

2 THE TRANSPRESSIVE SOUTHERN TERMINATION OF THE BUCARAMANGA

3 FAULT (COLOMBIA): INSIGHTS FROM GEOLOGICAL MAPPING, STRESS

4 TENSORS, AND FRACTAL ANALYSIS

PT
6 Francisco VELANDIA a,*b, Mauricio A. BERMÚDEZ c,d

RI
7

SC
a
9 Departamento de Geociencias, Universidad Nacional de Colombia, Cra. 20 Nº 45-03,

10 Bogotá, D.C., Colombia, favelandiap@unal.edu.co

11
U
AN
*b
12 Escuela de Geología, Universidad Industrial de Santander, Cra. 27 Calle 9,
M

13 Bucaramanga, Colombia, favelanp@uis.edu.co, +57 3125272971

14
D

c
15 Escuela de Ingeniería Geológica, Universidad Pedagógica y Tecnológica de Colombia,
TE

16 Facultad Seccional Sogamoso, Calle 4 Sur N° 15- 134, Sogamoso, Colombia,

17 maberce@gmail.com
EP

18
d
Facultad de Ciencias Naturales y Matemáticas, Universidad de Ibagué, Cra. 22 Calle
C

19
AC

20 67B, Ibagué, Colombia.

21

22

23 Keywords: transpression; restraining bend; domino structure; stress tensor; Bucaramanga

24 Fault; Colombia

25
ACCEPTED MANUSCRIPT
26

27 ABSTRACT

28 This study provides a kinematic model of the southern termination of the Bucaramanga

29 Fault (BF), in the Eastern Cordillera of Colombia, through the integration of geological

30 information, field geology, morphostructural mapping, deduced stress tensors, and fractal

PT
31 analysis. Minor Riedel-type faults are related to the main fault trace and to its southern

RI
32 termination in multiple structures and allow a determination of the maximum horizontal

33 stress (SHm) azimuth as between 105° to 140°, while stress tensors linked mainly with the

SC
34 southern ending faults show an SHm with an azimuth ranging from 108° to 169°. The

U
35 southern termination of the BF comprises a series of sub-parallel faults that reach lengths
AN
36 of up to 60 km, and stress tensors solutions confirm the prevalence of pure strike-slip

37 motion. This set of faults constitutes a transpressive system in a restraining bend (positive
M

38 flower structure), characterized by a “domino” style, that captured the Boyacá and Soapaga

39 faults, two of the inverted structures along the axial zone of the Eastern Cordillera. A
D

40 general sinistral displacement of 23 km is assumed along the principal BF, which is


TE

41 distributed along the southern positive flower structure with kinematics that change from

strike-slip in the centre to oblique-slip on the lateral faults. We performed fractal analysis
EP

42

43 and obtained a D-value of D = 1.42 ± 0.07 for the whole area, in contrast a value of D =
C

44 1.37 ± 0.08 for the northern part and D = 1.39 ± 0.12 for the southern part of the BF, and
AC

45 taking into account the standard deviations, our analysis corroborates the existence of

46 similar faulting mechanisms and fractionation.

47

48

49

50
ACCEPTED MANUSCRIPT
51

52 1. Introduction

53 The Bucaramanga Fault (BF) is a regional strike-slip structure in the core of the Colombian

54 Eastern Cordillera and is considered the western limit of the Maracaibo Block and

55 Santander Massif. Yet the southward continuity of the BF toward the Floresta Massif is not

PT
56 entirely considered as a strike-slip fault (Ujueta, 2003). There are data and models for the

RI
57 interpretation of strike-slip faults related to the BF in the Floresta Massif (Acosta et al.,

58 2007; Cediel et al., 2003; Kammer, 1999; Sarmiento, 2001; Taboada et al., 2000; Velandia,

SC
59 2005). However, most of the Eastern Cordillera models show thick and thin skin styles of

60 compressional tectonics towards the eastern border of the massif, with the Boyacá and

61
U
Soapaga faults as the main structures. Some of these models include compression and
AN
62 inverse tectonics, with an abundant presence of thrusts, backthrusts, duplexes, and passive
M

63 roofs (Colletta et al., 1990; Cooper et al., 1995; Dengo and Covey, 1993; Roeder and

64 Chamberlain, 1995; Tesón et al., 2013; Toro, 1990; Toro et al., 2004; Saylor et al., 2012a).
D

65
TE

66 The Neotectonic activity of the BF as a sinistral strike-slip structure is well documented

67 (Diederix et al., 2008, 2009; Ingeominas, 1997; Jiménez et al., 2015; Paris et al., 2000;
EP

68 París and Romero, 1994; Veloza et al., 2012). The Soapaga and Boyacá faults are both

identified as potentially active (Ingeominas, 1997). However, there is not enough


C

69
AC

70 information in the area about the relationship between these three faults, except for

71 regional structural descriptions (Toro, 1990; Velandia, 2005) and studies with a

72 sedimentological focus (Kammer and Sánchez, 2006; Saylor et al., 2012b).

73

74 This research presents morphological analysis and structural mapping of the southern

75 termination of the BF in the Floresta Massif, as well as stress analysis to explain the
ACCEPTED MANUSCRIPT
76 current brittle deformation in the area. We analysed regional geological data comprising

77 Colombian Geological Survey maps compiled by Gómez et al. (2015, 2015) (Fig. 1) that is

78 1:100,000 scale charts: H-12 and H-13 (Ward et al., 1977 a, b), 136-Málaga (Vargas et al.,

79 1976), 152-Soatá (Vargas et al., 1987), and 172-Paz de Río (Ulloa et al., 1998). We

80 complemented the structural framework obtained from the geological mapping by

PT
81 incorporating regional and field analysis, including measurements of slickensides as

RI
82 kinematic indicators in rock outcrops along primary and secondary roads. This

83 methodology made it possible to obtain a transpressive system model for the southern

SC
84 termination of the BF. As part of our work, we defined the faults related to this termination

85 and their different structural styles and illustrated the contrast between the Boyacá and

86
U
Soapaga faults. Additionally, we calculated the fractal dimension to show that different
AN
87 fault arrays across the study area exhibit fractal properties at different scales.
M

88

89 2. Geological setting
D

90 The study area is located in the axial zone of the Eastern Cordillera, where structures
TE

91 transition from a NE trend to NW one and involve the Floresta Massif’s northeast margin

92 and the southern extreme of the Santander Massif. Basement rocks outcrop along both
EP

93 massifs, together with Paleozoic and Jurassic sedimentary units in contact with Cretaceous

and Paleogene marine, transitional, and continental sequences (Fig. 1). The core of the
C

94
AC

95 Eastern Cordillera is composed of these massifs, which were part of the Santander High,

96 which separated Mesozoic basins (De Freitas et al., 1997; Sarmiento, 2001). The area is the

97 result of tectonic inversion of extensional structures and exhumation during the Cenozoic

98 (Fabre, 1983; Colletta et al., 1990; Dengo and Covey, 1993; Cooper et al., 1995; Roeder

99 and Chamberlain, 1995; Kammer and Sánchez, 2006; Horton et al., 2010; Saylor et al.,

100 2012b; Tesón et al., 2013). The Mesozoic rocks that border the Neoproterozoic and
ACCEPTED MANUSCRIPT
101 Paleozoic crystalline and sedimentary units in the convergence zone between the

102 Bucaramanga, Boyacá, and Soapaga faults (Fig. 1) probably covered the Santander High

103 during the Late Mesozoic prior to Cenozoic exhumation (Horton et al., 2010; Parra et al.,

104 2010), when transpressive deformation was predominant (Ramírez et al., 2012).

105 FIG. 1 HERE

PT
106 The crystalline basement (Gómez et al., 2015; Vargas et al., 1976; 1987) contains high-

RI
107 grade metamorphic rocks (Bucaramanga Gneiss and Orthogneiss) and units with lower

108 grade metamorphism like the Silgará Formation and Guaca – La Virgen Metamorphites (or

SC
109 Metamorphic Floresta unit). In addition, upper Paleozoic and Mesozoic sedimentary rocks

110 crop out, as do intrusive rocks from the Ordovician-Silurian and Jurassic (into the

111
U
Neoproterozoic and Paleozoic basement units), and Jurassic volcanic units. The Cretaceous
AN
112 marine sequence is represented by sandstones, siltstones, and limestones of the Rio Negro,
M

113 Tibú-Mercedes, Aguardiente, Capacho, La Luna, and Colón-Mito Juan formations.

114 Although the nomenclature of the Soapaga Fault is different southward, the Cenozoic
D

115 sedimentary sequence includes units associated with the Catatumbo basin nomenclature,
TE

116 such as the Barco, Cuervos, Mirador, and Carbonera formations, correlated in the

117 Cordillera with the Socha’s Sandstones, Socha´s Claystones, Picacho, and Concentración
EP

118 formations, respectively. Quaternary deposits are abundant in the zone, primarily over the

Cretaceous and Cenozoic units, and along the sector between the Boyacá and Soapaga
C

119
AC

120 faults (Fig. 1).

121

122 The interpretation of a transpressive system in the Eastern Cordillera that involves the

123 Santander and Floresta massifs is based on structural data, paleostress analysis and seismic

124 data (Kammer, 1996, 1999; Taboada et al., 2000). Sarmiento’s (2001) compilation

125 supports a transpressional tectonic inversion model for the Cordillera with higher intensity
ACCEPTED MANUSCRIPT
126 of deformation along the strike-slip BF. The inversion of normal faults, such as Boyacá

127 and Soapaga during the beginning of the Andean orogeny was also modelled for the

128 eastern boundary of the Floresta Massif with a thick and thin skin structural style (Saylor et

129 al., 2012a; Toro, 1990). Transverse faults have also been identified (Villar et al., 2017). A

130 horsetail array of faults toward the southern end of the BF shows how the faults affected

PT
131 the sedimentation of Cretaceous basins (Kammer and Sánchez, 2006).

RI
132

133 The southern part of the BF (from Bucaramanga to Ricaurte) was mapped at 1:25,000 and

SC
134 the morphotectonics were studied by the Colombian Geological Survey (Diederix et al.,

135 2008; Osorio et al., 2008); in particular, the characterization of Quaternary deposits was

136
U
carried out in detail; thus, erosional forms and deformation indices are associated with fault
AN
137 traces. Based on morphostructural mapping, Galvis et al. (2014) proposed a braided or
M

138 lenticular pattern, involving special pressure ridges, which controls the course of the

139 Chicamocha along the BF (suggesting a recent deformation).


D

140
TE

141 The current tectonic frame has been documented to show the relationship between South

142 America with the Nazca and Caribbean plates and the complex strain of the Northern
EP

143 Andes, including the Maracaibo Block, by using GPS data (Trenkamp et al., 2002), focal

mechanism and seismicity (Corredor, 2003; Colmenares and Zoback, 2003; Cortés and
C

144
AC

145 Angelier, 2005; Vargas and Mann, 2013; Egbue et al., 2014).

146

147 3. Methods

148 3.1. Geological mapping and delineation of faults

149 We performed regional analysis from geological charts and detailed mapping of the

150 geological faults by interpreting hillshade images derived from digital elevation models
ACCEPTED MANUSCRIPT
151 with 30- and 12.5-m resolution (NASA – Alaska Satellite Facility 2015). Additionally, we

152 used 1:30,000 and 1:40,000 aerial photographs. This mapping allowed us to define the

153 continuity of the main and minor faults and to identify the structural and morphological

154 features related to kinematics, such as ridges, depressions and offset drainages. We verified

155 the morphological features of faults during fieldwork and collected data from mesoscale

PT
156 kinematic indicators to carry out the stress tensor analysis. The construction of cross-

RI
157 sections helps to explain the structural styles in the area. Based on this morphological and

158 structural information, we propose a deformation model for the faults related to the

SC
159 southern termination of the BF.

160

161
U
The stress inversion requires a rigorous process that begins with field data collection,
AN
162 processing, and interpretation. Slickensides were measured on the fault planes, considering
M

163 the criteria of Petit (1987) and Doblas (1998) to define the kinematics, with priority given

164 to Riedel-type fractures. Some plane features were observed in detail to obtain
D

165 complementary information that supports homogeneous data, according to suggestions by


TE

166 Sperner and Zweigel (2010) and Hippolyte et al. (2012). This information consisted of

167 subjective criteria to evaluate quality considering: (i) weathering of outcrops, (ii) lithology,
EP

168 (iii) preservation and sharpness of slickensides, (iv) number and quality of kinematics

indicators, and (v) cross-cutting relationships of planes. Considering these aspects, quality
C

169
AC

170 percentages were assigned to each slickenside and fault plane measured in the outcrops:

171 60% (low confidence), 70-80% (reliable), and 90-100% (certain).

172

173 3.2. Stress tensor analysis

174 At least four fault planes are required to calculate a stress tensor. If the planes show

175 different orientations, so that they represent the four minimum parameters required, they
ACCEPTED MANUSCRIPT
176 are considered as suitables, these are: the three principal stress axes (σ1, σ2, and σ3) and the

177 stress ratio R (σ2-σ3/ σ1-σ3) (Delvaux and Sperner, 2003). It is crucial to define R’ because

178 it allows for direct determination of the associated stress regime according to the

179 relationship with R: extensional (R’ = R), strike-slip (R’ = 2 - R), or compressional (R’ = 2

180 + R); or in a more specific way according to its value: radial extensive (σ1 vertical, R’ = 0-

PT
181 0.25), pure extensive (σ1 vertical, R’ = 0.25-0.75), transtensive (σ1 or σ2 vertical, R’ =

RI
182 0.75-1.25), pure strike-slip (σ2 vertical, R’ = 1.25-1.75), transpressive (σ2 or σ3 vertical, R’

183 = 1.75-2.25), pure compressive (σ3 vertical, R’ = 2.25-.75), or radial compressive (σ1

SC
184 vertical, R’ = 2.75-3.00) (Delvaux et al., 1997; Delvaux and Sperner, 2003; Tripathy and

185 Saha, 2013). For the processing of fault slip data, Win-Tensor 5.8 software was used

186
U
(Delvaux and Sperner, 2003). The program includes an interactive process of stress tensor
AN
187 calculation and data separation to obtain good quality stress tensor solutions. In Win-
M

188 Tensor, the interactive graphical “Right Dihedron Method” is used to determine the

189 possible range of orientations σ1 and σ3. We used these results as a starting point for
D

190 iterative grid-search by the “Rotational Optimization” technique. The misfit function in
TE

191 “Rotational Optimization” makes it possible to minimize the angular deviation between the

192 observed and theoretical slip directions and maximizes the shear stress magnitude on the
EP

193 focal planes. It is possible to check the variations of stress tensors according to the planes

used by the software and to filter low confidence data in accordance with field
C

194
AC

195 observations (cross-cutting) or because of mechanical incompatibility according to the

196 Mohr circle and locations of planes in the fields of unstable domain (neoformed or

197 reactivated faults) or stable domain (non-slip fracture planes).

198

199 At the end of the processing of all the site data, only one or multiple tensors with different

200 stress regimes may result. The quality factor for each stress tensor (QRt) mainly depends
ACCEPTED MANUSCRIPT
201 on the number of data (n), their confidence level, and the number of data used as a

202 percentage of the total number of measurements from a given site, among others (Delvaux

203 and Sperner, 2003, Sperner and Zweigel, 2010). Although the relative age of the deduced

204 stress tensors is determined mainly from field observations of cross-cutting relationships

205 between faults, this is not always possible. In these cases, it is necessary to use other

PT
206 aspects for timing inferences, such as Mohr circle analysis (to observe the location of data

RI
207 in the neoformed or reactivated faults fields); rock age (where the kinematics indicator is

208 measured); quality, preservation and sharpness of the striated fault plane; associated

SC
209 minerals on the plane; and number of planes with different orientations. The software

210 makes it possible to represent the stress tensor as either an arrow or a beach ball (similar to

211 the symbol for focal mechanisms).


U
AN
212
M

213 3.3. Fractal analysis

214 Because the surface plates of plate tectonics evolve in time, geometrical incompatibilities
D

215 develop (Dewey, 1975); in these cases, a universal scaling law is imperative to understand
TE

216 the processes of the evolution of faults across geological time, and thus fractal models can

217 be applied to tectonic fragmentation (Sammis et al., 1987). This last is caused by complex
EP

218 interactions of plates or tectonic blocks. Our study area is an excellent natural laboratory

for understanding fragmentation and fault generation processes. The fault pattern geometry
C

219
AC

220 has fractal, or self-similar, behaviour (Allègre et al., 1982; King, 1983; Turcotte, 1986).

221 We base the fractal models on fragmentation involving all scales during fault growth with

222 repetitive subdivision of breakage patterns (King, 1983; Turcotte, 1986). At least two

223 scaling laws describe the spatial distribution of faults (clustering) or the fault size

224 distribution (Bour and Davy, 1999; Davy et al., 1990, 1992). Two relevant exponents are
ACCEPTED MANUSCRIPT
225 the fractal dimension (D), and the exponent (a) of the frequency-length distribution related

226 through a = D + 1 in the fragmentation models (King, 1983; Turcotte, 1986).

227

228 We use a family of fractal dimensions to characterize the geometrical properties of a d-

229 dimensional set of faults (in our case d = 2) following Renyi (1971) and Hentschel and

PT
230 Procaccia (1983). The generalized dimension of order n (Dn; Goltz, 1997) is defined by:

1 ∑ 


 = lim
 − 1
→ 

RI
SC
231 where r is the scale or the size of a d-dimensional box, pi, is the probability that a point of

232 the set belongs to box i, and index i runs over nonempty boxes i = 1,…, j. In the case where

U
233 n = 0, the previous equation is called the Kolmogorov capacity (or box-counting
AN
234 dimension). To obtain this measurement, small cells (squares for sets embedded in two

235 dimensions, cubes for sets embedded in three dimensions, etc.) are used to cover the set.
M

236 The dimension is then defined as (Barton and La Pointe, 1995; Giaquinta et al, 1999;
D

237 Turcotte, 1992):


 =  = lim
TE

1
  

where  is the number of boxes that contain at least one part of the data and (r) is the
EP

238

239 length of a box. The fractal dimension D-value was calculated by the least-squares method
C

240 using the log(N(r)) versus log(1/r) plot. Typical slope values vary between 1 and 2
AC

241 (Mandelbrot, 1982).

242

For each structural map, we plot the local exponent  =




243 versus the box size R, we

244 select all fractal dimension values that are parallel to a straight line of zero slope
ACCEPTED MANUSCRIPT
245 (constant), and calculate the standard deviation of all these points by the following

246 expression:

247

1

% &
! = " # $ − 

'(
% is the mean fractal dimension of
where DR is the fractal dimension for each box I and 

PT
248

249 these points.

RI
250

SC
251 The box-counting method presents serious setbacks that must be considered when trying to

252 provide estimates of D (Giaquinta et al., 1999). However, this method can be used by

U
253 minimizing the number of boxes used at every chosen length scale (Goltz, 1997; Robertson
AN
254 et al., 1995).

255
M

256 To capture variation in clustering properties of the faults array and to better evaluate the
D

257 evolution of the fractal properties we performed a fractal analysis over three different
TE

258 structural maps. We calculated the fractal box dimension for the structural map (original

259 and filtered at a scale of 1:500,000), the second map, at a scale of 1:300,000, and the third
EP

260 map, at a scale of 1:200,000. To avoid spurious or biased numeric results, the structural

261 map must contain fractures whose sizes range over at least three orders of magnitude
C

262 (Pérez-López et al., 2005).


AC

263

264 4. Results

265 4.1. Sinistral strike-slip displacement

266 The length of the BF is approximately 320 km, and through the geological mapping

267 analysis of the structure to the south of the Santander Massif and along the Floresta Massif,

268 it was possible to recognize some zones where it would be feasible to measure the sinistral
ACCEPTED MANUSCRIPT
269 displacement, even though the geologic units do not present homogeneous contacts

270 alongside the strike-slip fault (Fig. 1).

271

272 The Paleozoic metamorphic unit (Silgará Fm.) outcropping in Piedecuesta could be

273 correlated with those that crop out in the northern Bucaramanga, which would indicate a

PT
274 sinistral displacement of approximately 18 km (Fig. 1). Further south, the Jurassic intrusive

RI
275 contact with the Silgará Fm. is offset by approximately 20 km between the Perchiquez and

276 Manco rivers faults, without considering the elongated intrusive body that outcrops along

SC
277 the BF damage zone. We observe the greater 30-km strike-slip displacement along the

278 main trace of the BF by comparing the southern extremes of the Bucaramanga Gneiss unit,

279
U
north at Río Manco Fault and south at Ricaurte. With the average of these three data, a 23-
AN
280 km displacement is assumed for the BF along its well defined single corridor.
M

281

282 The estimated 23-km displacement along the principal BF is redistributed along the
D

283 branches of the BF. Los Micos Fault displaces the Paleozoic metamorphic units by 11-km
TE

284 (Fig. 1). The Chaguacá Fault’s strike-slip displacement is calculated to be about 5 km from

285 three approximate measures: (i) 3 km for the Paleozoic intrusive body (Del Real, 2013),
EP

286 (ii) 5 km for the contact of Silgará Fm. and the intrusive unit, and (iii) 7 km displacement

for the Jurassic sedimentary rocks (Girón Fm.) at the southern extreme of the Chaguacá
C

287
AC

288 Fault. Toro (1990) also estimated a 5-km sinistral strike-slip for this fault. It is interesting

289 that around Onzaga, where different faults converge, Cretaceous units barely show 1.2 km

290 of sinistral displacement, so we assume that folding has absorbed the remaining 3.8 km

291 along with the secondary dextral and sinistral faults, as presented by Galvis (2016). We

292 have not considered the vertical displacement along the southern transpressive system of
ACCEPTED MANUSCRIPT
293 the BF, which is probably partitioned along all the structures, including lateral thrusts

294 faults (Fig. 1).

295

296 4.2. Mapping of the Bucaramanga Fault at its southern termination

297 The present identification of morphological features along the main trace of the BF

PT
298 between Bucaramanga and Ricaurte (Figs. 1 and 2) is not as detailed as that of Osorio et al.

RI
299 (2008) or that reported in Quaternary deposits (Diederix et al., 2008, 2009); however, our

300 integrated map that lets us discriminate a Riedel-type pattern that confirms the left lateral

SC
301 kinematics of the BF and its effect on the deformation of the basement and sedimentary

302 sequences (Figs. 1 and 2). Most of the mapped faults control the drainage pattern, being

303
U
occupied by rivers, which are the major structures, or streams, which are the local faults.
AN
304 The observed kinematics correspond to strike-slip faulting models with minor associated
M

305 synthetic (R) and antithetic (R’) faults (Fig. 3), whose intersections form lenses as the fault

306 evolves and a corridor or damage zone develops. Inactive traces located further away from
D

307 the principal displacement zone (Naylor et al., 1986; Tchalenko, 1970; Woodcock and
TE

308 Fischer, 1986) are also observed.

309 FIG. 2 HERE


EP

310 FIG. 3 HERE

Although the minor faults show a complex structure, from a regional point of view the
C

311
AC

312 southern BF can be divided into two distinct zones (Fig. 3). The first is an 82-km long

313 single corridor from Bucaramanga up to Ricaurte. The second extends 60 km southward

314 from Ricaurte to its intersection with the Soapaga Fault and is composed of several faults

315 that branch away from the main structure (Figs. 2 and 3). The northern corridor can be

316 subdivided into three sections (Fig. 4), which are named here, avoiding the repetition of

317 already known denominations for other faults. The Piedecuesta Section is 32 km long from
ACCEPTED MANUSCRIPT
318 Bucaramanga to the Río Manco transverse fault and defines the clear single corridor of the

319 main structure. In the 31-km-long Cepitá Section, bounded by the Río Manco and Guaca

320 faults to the north and south, respectively, the fault is more complex. The northern part of

321 the Cepitá’s section corresponds to the “Pescadero” descent to the Chicamocha River,

322 while the southern part of the fault zone crosses the mountainous area toward the river's

PT
323 right-hand margin. Through the Cepitá section, the BF seems to exhume part of the brittle-

RI
324 ductile crust zone given the outcrops of identified corresponding structures (Villamizar,

325 2014) and presence of pseudotachylites (González and Jiménez, 2014). The 19-km-long

SC
326 Ricaurte Section contains synthetic Riedel faults (R) that control the course of the

327 Chicamocha River. Galvis et al. (2014) mapped transpressive and transtensive landforms

328
U
that exhibit a lenticular or braided pattern between Cepitá and Ricaurte. With the exception
AN
329 of a small portion of the Piedecuesta Section near the city of Bucaramanga, where eight
M

330 Holocene earthquakes ranking from 7.0 to 7.4 Ms have been described (Diederix et al.,

331 2009), there is little paleoseismological characterization of the faults, and thus each section
D

332 cannot be defined as a segment. Nevertheless it is expected that under the current stress
TE

333 field, large strike-slip faults may produce movements at larger recurrent intervals.

334 FIG. 4 HERE


EP

335 To determine the simplicity or complexity of these sections and the southern terminal zone

of the BF, we mapped traces, geomorphic evidence, and kinematic indicators along the
C

336
AC

337 faults, such as saddles, benches, triangular facets, ridges, basins, offset drainage, trenches,

338 and hook-shapes (L-shaped spurs), among others (Fig. 2). With fewer indicators, the fault

339 is simpler along the Piedecuesta Section (Fig. 2a). In contrast, the Ricaurte section contains

340 numerous features such as pressure ridges (Fig. 2b). Mapping indicates that the ridges are

341 formed related to local restraining bends and between the main fault trace and the

342 secondary synthetic faults (R). Some of the ridges are large enough to extend up to 8 km in
ACCEPTED MANUSCRIPT
343 length, like the one that shutters the Chicamocha River course around Ricaurte (Figs. 2b, 3,

344 and 4) and present inner transverse faults as conjugate structures similar to the regional

345 pattern that is observed in the area, which is explained later.

346

347 From the general fault array and the associated Riedel traces, it is possible to obtain the

PT
348 local stress tensor orientation with the corresponding directions of principal stress axes

RI
349 matching a sinistral strike-slip model (Woodcock and Shubert, 1994, in Davis et al., 2012)

350 (Figs. 2 and 3). A WNW – ESE (approximately 105°) orientation for σ1 is obtained for the

SC
351 northern part of the fault, while the faults related to the southern termination take a more

352 northward strike and the stress tensor changes to NW – SE with an orientation of 140° for

353
U
σ1. The continuation of the Riedel pattern is reflected in the sinistral displacements within
AN
354 the Floresta Massif and across the Soapaga Fault (Figs. 2c,d and 4). Rose diagrams of the
M

355 mapped faults show the orientation distribution of the strands for each window (Fig. 2a-d)

356 confirming the angular relationship of the Riedel pattern and the main fault trace.
D

357
TE

358 From fault mapping, we deduce that the structure is neither simple nor narrow; instead, it

359 presents a broad (5 to 8 km) damage zone extending between Bucaramanga and Ricaurte
EP

360 (Fig. 4). We also identify five branched faults, related to the southern termination of the

BF, and at least three of them show strike-slip features.


C

361
AC

362

363 This general model (Fig. 3) is compared with some of the regional strike-slip faults

364 compiled by Mann (2007) in a global catalogue and the termination of the BF could

365 correspond to another example of a restraining bend in a sort of lazy z shape (Figs. 3 and

366 4), also similar to the restraining stepovers obtained in analogous modelling of strike-slip

367 faults (McClay and Bonora, 2001), taking into account that the wide single corridor
ACCEPTED MANUSCRIPT
368 represents the main trace along which the main strike-slip displacements occur, while in

369 the regional restraining bend, strike deformation is minor and accommodated better by

370 shortening (Fig. 1). The presence of different type of ridges along the BF (Figs. 2 and 3)

371 can be explained as result of local restraining bends, some under the influence of oblique

372 structures, examples of which are presented by Mann (2007). The longitudinal and internal

PT
373 faults of the regional restraining bend show a pattern of mixed-type of damage zones along

RI
374 a strike-slip fault, comparable to some cases cited by Kim et al. (2004) at different scales.

375

SC
376 4.3. Stress tensor analysis

377 Kinematic indicators, especially slickensides, were considered for different locations in the

378
U
area. Stress tensor analysis has been used to find the variation of the stress field across time
AN
379 and space (Angelier, 1994), with data from not only Quaternary deposits but also ancient
M

380 rocks in complex geological contexts, where a similar orientation of the main

381 compressional stress can be assumed as a solution of the regional tensor (Sassi and Faure,
D

382 1996; Tripathy and Saha, 2013). Similar studies of stress tensors have been developed
TE

383 along the Eastern Cordillera of Colombia and its borders (Kammer, 1999; Taboada et al.,

384 2000; Cortés et al., 2005).


EP

385

We selected 30 field stations according to stress tensor quality (Fig. 4 for the locations) and
C

386
AC

387 prioritized those along the southern termination of the BF. See Table 1 for the stress

388 tensors obtained from the measured fault slip data. We measured a total of 758 slickensides

389 along the 30 sites, of which 678 (87%) were used as subsets to define 30 stress tensors.

390 Most planes were taken in intrusive igneous rocks (16) and competent sedimentary rocks

391 (nine in sandstones and one in conglomerates), while only four sites correspond to

392 metamorphic rocks from the Paleozoic (Table 1). For this analysis, only one stress tensor
ACCEPTED MANUSCRIPT
393 per site was considered; in most of the cases (26 of 30) it corresponded to the unique or the

394 first output stress tensor once the software (Win-Tensor 5.8) had processed data from each

395 site.

396 TABLE 1 HERE

397 Stress tensors indicate a NW – SE general orientation for the maximum horizontal stress,

PT
398 which varies between 108° and 169°, showing more parallelism in those sites located near

RI
399 the strike-slip main trace of the BF and more obliqueness in sites along the restraining

400 bend (Fig. 5). The obtained regimes imply mainly pure strike-slip, with some oblique

SC
401 variations to extension or compression, although there are a couple of purely extensional

402 stress tensors (6 and 22) east of the main Bucaramanga structure (Table 1, Figs. 4 and 5).

403
U
AN
404 Figure 4 shows the stereographic projections in the lower hemisphere of the fault planes
M

405 used to determine the stress tensors for each field site and also displays the corresponding

406 striae. Most stress tensors include planes with at least two directions, an aspect that can be
D

407 considered as complementary to the software’s QRt quality criteria, which are
TE

408 predominately based on the number of data per subset and slip deviation (α), although the

409 experience of the user and type of data (in this case only slickensides) are also considered.
EP

410 The orientation of the principal stress axes indicates that in at least half of the sites, two

axes remain sub-horizontal and the other is vertical, which implies better resolved stress
C

411
AC

412 tensors, like those where σ2 appears vertical, indicating a pure strike-slip regime, or the few

413 where σ1 is almost vertical, representing an extensive regime. Stress tensors with axes

414 gentle to moderate plunging axes can be related to sites in blocks with some tilting.

415 FIG. 5 HERE

416 Mohr circles help constrain the stress regime that each tensor represents according to the

417 relative principal stresses along N (normal stress) as well as the mechanical compatibility
ACCEPTED MANUSCRIPT
418 for each of the 30 stress tensors (Fig. 6). This analysis was guided to make certain that

419 planes were located in the neoformed faults field, avoiding data in the stable field, which

420 would belong to different, likely older stress tensors.

421

422 The stress tensor representation is facilitated by the Win-Tensor program (Delvaux and

PT
423 Sperner, 2003) in lower hemisphere projections (beach balls), similar to focal mechanism

RI
424 symbols (Fig. 5), where dark quadrants are related to compression and white quadrants to

425 extension. The analysis of the beach balls enables a visual comparison of the relationship

SC
426 between the mapped fault orientations and those determined by the software. Indeed, most

427 of the stress tensors show sub-vertical strike-slip faults with a N-S or NNW orientation,

428
U
which correspond to the trend of the regional geologic fault around the site. Some stress
AN
429 tensors involve planes that can be better related to synthetic (R) or antithetic (R’) faults. In
M

430 all the cases, the stress tensors present a NW-SE orientation (varying to WNW - ESE),

431 which induces sinistral strike-slip displacements along the main longitudinal faults or
D

432 along synthetic Riedel faults (R) (Figs. 2, 4, and 5).


TE

433 FIG. 6 HERE (Part 1 and Part 2)

434 4.4. Southern transpressive system of the Bucaramanga Fault


EP

435 We present a model to explain the kinematics and structural style of the faults related to the

BF’s southern termination (Fig. 7). The model is the result of geological and
C

436
AC

437 morphostructural mapping analysis, and linked stress tensors as consequence of the current

438 stress field.

439

440 A set of faults along the Chicamocha River is identified as related to the BF around the

441 Ricaurte site (Figs. 5 and 7). The Chaguacá Fault remains more similar to the BF regarding

442 west vergence. This fault has also been referred to as the Onzaga Fault (the trace north
ACCEPTED MANUSCRIPT
443 from the town towards the Chicamocha River); however, we recommend abandoning the

444 use of the name “Onzaga Fault” to avoid confusion, since at least four faults converge

445 around that locality (Fig. 4). The 60-km-long Chaguacá Fault continues southward to cut

446 the Soapaga Fault south of Paz de Río town (Fig. 7).

447

PT
448 To the east, and parallel to the Chaguacá Fault, we can follow the Los Micos Fault (Galvis,

RI
449 2016). It is 58 km long and has a sense of motion to the east (opposite vergence to that of

450 the BF). It has been previously identified by the term “principal structure” by Vargas et al.

SC
451 (1987) and Toro (1990). This Los Micos Fault also cuts the Soapaga Fault north of Paz de

452 Río.

453
U
AN
454 Another parallel and longitudinal structure is La Chorrera Fault (Fig. 7), which has a length
M

455 of 40 km; it arises in the north part of Ricaurte, and along its path it separates the Paleozoic

456 rock in the west from the Cretaceous rocks in the east. From this, in turn, the Chicamocha
D

457 Fault branches off affecting the Cretaceous sedimentary sequence and continues southward
TE

458 up to join the Soapaga Fault.

459 FIG. 7 HERE


EP

460 West of the Chaguacá Fault, El Topón Fault is exposed with a length of 30 km;

morphostructural features along the fault still define sinistral strike-slip kinematics for it
C

461
AC

462 (Fig. 2d). Southward, the Chaguacá Fault joints the Pargua Fault, which is 52 long from

463 Onzaga up to the Soapaga Fault. The Pargua is interpreted as a west-verging reverse fault,

464 and constitutes the western border of the southern transpressive system of the BF.

465

466 Inside this transpressive system and bounded by the Chaguacá, Los Micos, and La

467 Chorrera longitudinal faults, secondary and transverse faults are present with NE or ENE
ACCEPTED MANUSCRIPT
468 trends, most of them with dextral displacement, opposite to the sinistral movement of the

469 principal faults. This array is repeated at more detailed scales, as is interpreted from some

470 maps like Vargas et al. (1987) (Fig. 7b). The cross-sections (Fig. 8), transverse to the

471 longitudinal faults (Fig. 7), show a symmetric positive flower structure for the

472 transpressive system (regional restraining bend), with Chaguacá, Los Micos and La

PT
473 Chorrera as the principal strike-slip faults, from which lateral faults arise with a greater

RI
474 reverse component and more vergence to the east (El Olivo and Chicamocha faults) and

475 west (Pargua Fault). The faults located to the west, like Pueblo Viejo and Boyacá (Figs. 1

SC
476 and 7), do not form part of the flower structure, but are instead considered as structures of

477 inversion tectonics, as discussed later.

478
U
FIG. 8 HERE
AN
479 4.5. Fractal analysis
M

480 A fractal analysis is carried out since the same (self-similar) behaviour is observed in the

481 structural pattern of the study area at different scales, as is proposed for the regional brittle
D

482 deformation of the Santander Massif (Velandia, 2017). Fractal dimensions for transcurrent
TE

483 fault arrays lie between 1.4 and 1.6 (Guarnieri and Vinciguerra, 2004). The analysis of the

484 BF array was achieved by performing different box-counting measurements to evaluate the
EP

485 extent to which this fault array might be fractal. Also with this analysis, it was possible to

investigate the geometry relative to the fracturing processes of different deformational


C

486
AC

487 events, and we divided the study area into four distinct sectors at different scales (Fig. 9).

488 Table 2 shows the D-value of every single sector: each has a fractal dimension of

489 approximately 1.4.

490

491 The values obtained here for the fractal dimension of the spatial fault distribution indicate

492 that the complexity of the 1:500,000 structural map (D = 1.41) is larger than that of the
ACCEPTED MANUSCRIPT
493 1:200,000 maps; specifically, D = 1.36 for the northern map, and D = 1.39 for the southern

494 map. A higher dimension means that the full map includes a greater level of spatial fault

495 distribution in comparison with the smaller scale map (Figs. 10 and 11). However, this

496 property is evident because the complete map includes fractures generated by different

497 tectonic events. Even so, the comparison between the fractal dimensions obtained for maps

PT
498 at smaller scales is interesting because these maps represent the fracture pattern due to a

RI
499 particular modern stress field. Furthermore, this result corresponds to the property of

500 fractal sets, whereby the fractal dimension of the set included within another set is either

SC
501 less than or equal to the embedding set (Hasting and Sugihara, 1993). In the fractal

502 analysis for the 1: 500,000 scale map the longest lineament is 90 km, and the shortest is 0.6

503
U
km. The size of the different main fault lineations, with their minimum and maximum
AN
504 lengths, where found through a routine added into ArcGIS to calculate lengths of linear
M

505 features; however, this method is not shown here because it is beyond the scope of this

506 paper (see Abdullah et al., 2013). Additionally, the lineaments pattern is representative of
D

507 the strain accommodation from different tectonics events, as shown by Scholz and Cowie
TE

508 (1990), and Marret and Allmendinger (1991). As suggested by Pérez López et al. (2005),

509 the lower limit of the box size is obtained from the length of the smallest fault, and the
EP

510 upper value corresponds to the size of the maximum blank space on the different structural

maps.
C

511
AC

512 FIG. 9 HERE

513 TABLE 2 HERE

514 The fractal dimension of smaller scale maps indicates the degree of complexity due to the

515 geometric irregularity of spatial fracture patterns caused by the different stress fields. As

516 the fractal dimension is larger on the southern map (D = 1.39) than on the northern one (D

517 = 1.36), the pattern of the Soapaga and Boyacá faults is more complex than that of the
ACCEPTED MANUSCRIPT
518 northern part. Because the southern pattern is shaped by newly formed and several

519 reactivated families of fractures (NW–SE), the spatial complexity is greater than that in the

520 northern stress field. This observation is in accordance with the fractal analysis carried out

521 above.

522 FIG. 10 HERE

PT
523 FIG. 11 HERE

RI
524 5. Discussion

525 5.1. Sinistral displacement along the Bucaramanga Fault

SC
526 Large amounts of displacement that were initially attributed to the named Santa Marta –

527 Bucaramanga Fault have already been questioned by Toro (1990) and Ujueta (2003). They

528
U
cited the works of Campbell (1968) and Irving (1971) to indicate that their estimates
AN
529 lacked valid piercing points to calculate the proposed displacements of 110 km. Toro
M

530 (1990) considers only 45 km of sinistral slip along the BF, which he distributes southward,

531 along thrusts of the Boyacá, Soapaga-Chicamocha and Guantiva faults. We propose 23 km
D

532 of strike-slip displacement for the main trace of the BF, similarly redistributed along the
TE

533 longitudinal faults that make up part of the southern transpressive flower structure or

534 restraining bend, where the offset is lower than that along the northern main trace (Figs. 1
EP

535 and 3). The difference in displacement with respect to Toro (1990) can be partially

explained by compression, that is, thrusting, along the Boyacá and Soapaga faults prior to
C

536
AC

537 their capture by the BF.

538

539 5.2. Structural model for the southern termination of the BF

540 The almost parallel longitudinal faults that make up part of the southern termination of the

541 BF, and theirs sinistral strike-slip kinematics (as derived by morphostructural features and

542 NW - SE maximum principal compressive stress), in conjunction with the inner dextral
ACCEPTED MANUSCRIPT
543 faults, allow us to propose a model similar to a strike-slip domino structure (bookshelf

544 mechanism) for the brittle deformation in the area (Fig. 7), implying that the faults show

545 not parallelism but a lenticular pattern, like those presented by Sanderson et al. (1991) for

546 cases in west Spain and along the San Andreas Fault in California. Although the domino

547 structures are most known for normal faults and microscopic fabrics, some analogues have

PT
548 been proposed for strike-slip rhomboidal geometries affecting the basement in the NE

RI
549 Atlantic margin (Doré et al., 1997) and examples shown in California (Christie-Blick and

550 Biddle, 1985), Spain (Sanderson et al., 1991), the Ural region (Kolodyazhnyi, 2015), and

SC
551 Turkey (Koç and Kaymakci, 2013). Mandl (1987) mentioned the tectonic importance of

552 the bookshelf mechanism that operates in strike-slip zones, citing some examples in

553
U
Venezuela, France, and Saudi Arabia. Here, the model is also corroborated by local
AN
554 structures at different scales such that it suggests that the southern termination of the BF, in
M

555 the Floresta Massif, exhibits a fractal domino pattern (Figs. 7 and 9), taking into account

556 that this domino behaviour in domino is interpreted in regional structural models for the
D

557 Santander Massif (Restrepo-Pace, 1995; Velandia, 2017). Regarding these types of similar
TE

558 geometries across a wide ranges of scales, Kim et al. (2004) present different models and

559 examples of fault damage zones, including strike-slip structures with rotated blocks in
EP

560 domino patterns.


C

561
AC

562 5.3. The southern termination of the BF within the axial zone of the Eastern Cordillera

563 Concerning the evolution of the main fault structures to the axial zone and NE of the

564 Eastern Cordillera, the BF is most likely linked to the development of the Boyacá and

565 Soapaga faults to the south since at least the early Mesozoic, as deduced by Kammer and

566 Sánchez (2006). Inversion tectonics of normal faults in the Floresta Massif were also

567 dominated by sinistral displacement along the BF. The results of our study determine that
ACCEPTED MANUSCRIPT
568 the southern termination of the Floresta Massif is defined by the transpressive structure

569 with a predominantly strike-slip movement along the principal faults and thrusting along

570 the lateral faults, resulting in a positive flower structure (Figs. 7 and 8) similar to regional

571 restraining bends and stepovers along strike-slip fault systems (McClay and Bonora, 2001;

572 Mann, 2007). The main strike-slip trace of the BF maintains the general NNW trend of the

PT
573 Eastern Cordillera, while the restraining bend of the BF termination represents a transition

RI
574 zone with the NE trending of the Cordillera, including structures such as the Boyacá and

575 Soapaga faults along its axial zone.

SC
576

577 The brittle deformation model presented here suggests that the BF captures the Boyacá

578
U
Fault in the Onzaga area (Figs. 7 and 8b). South of Onzaga, however, the faults have
AN
579 different paths and structural evolution. Boyacá is primarily a reverse fault, while the BF
M

580 continues as a strike-slip fault with a clear positive flower structure. For this reason, our

581 interpretation is that the BF does not end as a horsetail structure. Another reason for this
D

582 analysis is that south of the fault’s termination, the kinematics of strike-slip changes from
TE

583 sinistral (BF) to dextral along the ancient and reactivated Boyacá and Soapaga faults

584 (Kammer, 1999; Velandia, 2005).


EP

585

The Pueblo Viejo Fault (Fig. 7) constitutes a backthrust of the Boyacá Fault, and both
C

586
AC

587 define the most meridional outcrop of the basement of the Eastern Cordillera in the

588 Floresta Massif. Probably, before inversion, the Pueblo Viejo Fault was part of the

589 Mesozoic basin’s border, which would have started with the deposit of the volcanic and

590 volcaniclastic rocks observed along and between these two faults, as part of the dextral

591 horsetail structure mentioned by Kammer and Sánchez (2006) for the early Mesozoic. In

592 the present model, the Pueblo Viejo Fault was also captured by the Chaguacá Fault
ACCEPTED MANUSCRIPT
593 between Ricaurte and San Joaquín (Fig. 7), most probably during the tectonic inversion,

594 accelerated by activity on the BF.

595

596 To the south, the relationship of the BF with the Soapaga Fault seems to be more recent, as

597 mentioned by Toro (1990), who referred to the continuity of the Soapaga Fault with the

PT
598 Chicamocha Fault. Here, we show that the Chicamocha Fault constitutes the eastern border

RI
599 of the flower structure and its movement takes place mainly along thrusts with east

600 vergence (Figs. 7 and 8b), since in the regional restraining bend the longitudinal strike-slip

SC
601 faults are located to the centre, and thrust faults to the flanks. The capture of the Soapaga

602 Fault occurs where the strike changes from N - S to SSW (Fig. 7). Southward from this

603
U
place, the Soapaga Fault is cut at another three sites by the transpressive structure: north
AN
604 and south of Paz de Río by Los Micos and Chaguacá faults, respectively, and further south
M

605 by El Topón Fault.

606
D

607 The similarities between the Soapaga-Chicamocha and Río Servita faults (Fig. 8a) are also
TE

608 manifested by their structural styles and by the geological units outcropping sideward to

609 the faults, which indicates that these faults could have constituted one large regional
EP

610 structure (Velandia, 2017) from the time of their normal kinematics as part of the

Mesozoic basin, currently inverted along the axial zone of the Eastern Cordillera.
C

611
AC

612

613 5.4. Fractal dimension, stress tensors, structural model and the current tectonic

614 configuration of the area

615 Firstly, it is appropriated to check the range in which the fault array shows fractal

616 properties. A fractal dimension of D = 1.42 ± 0.07 was obtained for the entire study area;

617 in contrast, values of D = 1.37 ± 0.08 and D = 1.39 ± 0.12 were obtained for the northern
ACCEPTED MANUSCRIPT
618 and southern parts of the BF, respectively. These three sectors are affected by a similar

619 tectonic process. In order to investigate the geometry relative to the fracturing processes of

620 the different deformational events, we considered the spatial distribution of faults and the

621 lithotypes outcropping in the study area. For the entire study area, the D-values obtained

622 suggest a superimposition of the transpressive structure and the strike slip-faults are

PT
623 described by the highest fractal value obtained (D = 1.42 ± 0.07). The southern part (D =

RI
624 1.39 ± 0.12) describes the geometry of the strike-slip fault; however a high value of the

625 standard deviation suggests a major presence of subsidiary strike-slip faults bounded by

SC
626 two systems of transverse faults. The result (D = 1.37 ± 0.08) for the northern part of the

627 BF is evidence of the powerful fractal analysis. Apparently, the different thrust

628
U
outcroppings absorb a huge part of the deformation. Our results are comparable with
AN
629 values obtained for the northwestern Sicily Neogene fault system (Guarnieri et al., 2004).
M

630 The deformation regime today is evolving as a transpressive system in a homogeneous

631 way, with greater compression and reactivation of faults towards the south. This is
D

632 compatible with the stress tensor analysis.


TE

633

634 The orientation of the stress tensor changing from WNW - ESE to NW - SE in the area
EP

635 coincides with those shown previously for this part of the Eastern Cordillera (Floresta

Massif) by Kammer (1999) and Taboada et al. (2000) through observation of fault planes
C

636
AC

637 and with the present-day principal stress direction indicated by Egbue et al. (2014) from

638 borehole breakouts, focal mechanisms, and GPS data, related to the collision of Panamá

639 and the Northern Andes. Although the NW - SE stress tensor is persistent in the Cordillera,

640 some other W - E and WSW - ENE orientations have been observed in Quaternary

641 deposits in Bucaramanga (Taboada, et al., 2000), more closely related with the

642 northeastward escape of the Northern Andes (Egbue et al., 2014).


ACCEPTED MANUSCRIPT
643

644 The NW - SE stress tensors and the morphostructural mapping obtained here confirm the

645 sinistral strike-slip kinematics that Velandia (2005) calculated in a general way up to 5°50’

646 latitude, but the transpressive model proposed here differs from the duplex structure he

647 proposed earlier. This because the observed dextral displacements along the secondary

PT
648 transverse faults do not correspond to the movement along inner faults of a duplex

RI
649 structure, which is supposed to have the same sinistral kinematic as the main longitudinal

650 faults (Woodcock and Fischer, 1986). By contrast, the structural style observed for the

SC
651 almost parallel and sinistral regional faults, together with the dextral displacement of the

652 NE minor faults inside the regional structure, better corresponds to a strike-slip domino

653
U
model, like those observed in areas where strike-slip faults dominate (Christie-Blick and
AN
654 Biddle, 1985; Mandl, 1987; Sanderson et al., 1991; Koç and Kaymakci, 2013;
M

655 Kolodyazhnyi, 2015), but in a sort of lenticular pattern, and even across a wide range of

656 scales (Kim et al., 2004). The fractal nature of the domino pattern means that it is repeated
D

657 at different scales and mainly along the regional longitudinal faults as in the Santander
TE

658 Massif (Velandia, 2017).

659
EP

660 Recent sinistral strike-slip displacement along the BF has been reported by París et al.

(2000) for the Quaternary alluvial fan (at the city), since its apex is cut and separated 2.5
C

661
AC

662 km from its source, the Suratá river (Diederix et al., 2008; Jiménez et al., 2015). We find

663 little evidence of Neotectonic activity since no Quaternary deposits are cut by any of the

664 faults along the southern termination of the BF. However, there is at least one local and

665 elongated morphological depression, apparently controlled by the Chaguacá Fault (Del

666 Real, 2013), that is filled with recent sediment (Figs. 1 and 2d). Galvis (2016) proposed

667 possible damming of the Chicamocha River and the formation of an extensive lake during
ACCEPTED MANUSCRIPT
668 the Pleistocene due to neotectonic activity related to the huge shutter ridge (El Tendido)

669 located in front of Ricaurte (Figs. 2b, 3, and 4). The lake and its age were reported by

670 Villarroel et al. (2001) around Soatá town, in a sector located 35 km upstream of the

671 shutter ridge.

672

PT
673 6. Conclusions

RI
674 A general sinistral strike-slip displacement of 23 km is assumed along the main BF, with

675 southward partitioning along the transpressive termination faults, either as pure strike-slip

SC
676 along Los Micos and Chaguacá faults or with a reverse component along lateral faults of

677 the flower structure.

678
U
AN
679 Morphostructural mapping of the BF allows us to observe a single and wide corridor with a
M

680 length of 82 km between Bucaramanga and Ricaurte. Complex and involving ridges and

681 elongated valleys that are controlled by synthetic (R) and antithetic (R’) Riedel faults
D

682 confirm the sinistral kinematics and determine a damage zone varying from 5 to 8 km
TE

683 wide. Based on this mapping, we subdivided this part of the BF into three sections:

684 Piedecuesta (32 km length), Cepitá (31 km) and Ricaurte (19 km).
EP

685

The analysis of 30 stress tensors at the same number of sites show a general NW - SE
C

686
AC

687 direction for the maximum horizontal stress (σ1), wjose azimuth varies from about 108° to

688 169° azimuth, under predominant pure strike-slip stress regimes. Most stress tensors show

689 sub-vertical strike-slip faults oriented N - S or NNW, validating the regional geological

690 fault at each site or in the surroundings. In all the cases, the stress tensors present a NW -

691 SE orientation, which induces sinistral displacements along the principal longitudinal

692 faults and subparallel synthetic (R) traces.


ACCEPTED MANUSCRIPT
693

694 Geological and geomorphological mapping, along with stress tensors analysis, permits us

695 to conclude that the BF involves a southern termination, 60 km length, in the Floresta

696 Massif, with sinistral and sub-parallel faults, and its structural style corresponds to a

697 transpressive system (restraining bend) in a domino pattern (lenticular) with fractal

PT
698 behaviour.

RI
699

700 The southern transpressive system of the BF, while developing southward, has captured the

SC
701 Boyacá and Soapaga faults, two of the inverted structures along the axial zone of the

702 Eastern Cordillera. The Soapaga Fault is cut off in several places accordingly as it has been

703
U
reached by the associated longitudinal strike-slip faults.
AN
704
M

705 The values of the fractal dimension obtained for the different scales of structural maps of

706 the southern termination of the BF made it possible to compare deformational events. The
D

707 obtained results suggest that: (i) the fractal dimension computed for the whole fault array
TE

708 (D =1 .41) is consistent with a fractal geometry of strike-slip faulting; (ii) the faulting

709 processes for the northern (D = 1.36) and southern (D = 1.39) BF suggest a current similar
EP

710 mechanism of faulting.


C

711
AC

712 Acknowledgements

713 This work is part of the doctoral thesis of Francisco Velandia at the Universidad Nacional

714 de Colombia, supported by Colciencias-Colfuturo grant (PDBC 6172) and the study

715 commission given by the Universidad Industrial de Santander. The authors are especially

716 grateful for discussions with Alfredo Taboada about stress tensors concepts, as well as for

717 reading of the manuscript by Greg Hoke, Carlos Zuluaga, Jorge Villalobos and David
ACCEPTED MANUSCRIPT
718 Buhagiar. We thank Diego Osorio, Helbert García, and Jorge Baquero for their assistance

719 during the fieldwork. The Universidad de Ibagué (Project 15-377-INT to MAB) is also

720 thanked for time to Mauricio Bermúdez. We are very grateful to the reviewers Paul Mann

721 and Dilip Saha for their useful remarks and contributions.

722

PT
723 References
724

RI
725 Abdullah, A., Nassr, S., and Ghaleeb, A., 2013. Remote sensing and geographic
726 information system for fault segments mapping a study from Taiz area, Yemen. Journal of
727 Geological Research, Article ID 201757. https://doi.org/10.1155/2013/201757.

SC
728
729 Acosta, J., Velandia, F., Osorio, J., Lonergan, L., Mora, H., 2007. Strike-slip deformation
730 within the Colombian Andes. Deformation of the Continental Crust. Geological Society of
731 London, Special Publications 272, 303–319.

U
732
733 Allègre, C.J., Lemouel, J.L., Provost, A., 1982. Scaling rules in rock fracture and possible
AN
734 implications for earthquake prediction. Nature 297, 47–49.
735
736 Angelier, J., 1994. Fault slip analysis and palaeostress reconstruction. In P. L. Hancock
M

737 (Ed.), Continental Deformation. Oxford: Pergamon Press, 53–100.


738
739 Audemard, F.E., Audemard, F.A., 2002. Structure of the Mérida Andes, Venezuela:
D

740 Relations with the South America‐Caribbean geodynamic interaction. Tectonophysics 345,
741 299–327.
TE

742
743 Barton, C., La Pointe, P.R., 1995. Fractal analysis of scaling and spatial clustering of
744 fractures. In: Barton, C., La Pointe, P.R. (Eds.), Fractal in Earth Science. Plenum Press,
745 New York, 141–178.
EP

746
747 Bour, O., Davy, P., 1999. Clustering and size distributions of fault patterns: theory and
748 measurements. Geophysical Research Letters 26 (13), 2001–2004.
C

749
750 Campbell, C.J., 1968. The Santa Marta wrench fault of Colombia: Transactions, 4th
AC

751 Caribbean Geological Conference, Port-of-Spain, Trinidad, 247–261.


752
753 Cediel, F., Shaw, R., Cáceres, C., 2003. Tectonic assembly of the northern Andean Block.
754 In: Bartolini, C., Buffler, R. T., and Blickwede, J. (Eds.), The Circum-Gulf of Mexico and
755 the Caribbean: Hydrocarbon habitats, basin formation and plate tectonics. American
756 Association of Petrolum Geologists Memoir 79, 815–848.
757
758 Christie-Blick, N., Biddle, K., 1985. Deformation and basis formations along strike-slip
759 faults. In: Biddle, K. and Christie-Blick, N. (Eds.), Strike-slip deformation, basin
760 formation, and sedimentation. Society of Economic Palaeontologists and Mineralogistics,
761 Special Publication 37, 1–4.
762
ACCEPTED MANUSCRIPT
763 Colletta, B., Hebrard, F., Letouzey, J., Werner, P., Rudkiewicz, J., 1990. Tectonic style and
764 crustal structure of the Eastern Cordillera (Colombia) from a balanced cross-section. In:
765 Letouzey, J. (Ed.), Petroleum and tectonics in mobile belts: Paris, Editions Technip, 81–
766 100.
767
768 Colmenares, L., Zoback, M., 2003. Stress field and seismotectonics of northern South
769 America. Geology 31: 721-724.
770
771 Cooper, M., Addison, F., Alvarez, R., Coral, M., Graham, R., Hayward, A., Howe, S.,

PT
772 Martínez, J., Naar, J., Peñas, R., Pulham, J., Taborda, A., 1995. Basin development and
773 tectonic history of the Llanos Basin, Eastern Cordillera, and Middle Magdalena Valley,
774 Colombia. American Association of Petroleum Geologists Bulletin 79 (10), 1421–1443.

RI
775
776 Corredor, F., 2003. Seismic strain rates and distributed continental deformation in the
777 northern Andes and three-dimensional seismotectonics of the northwestern South America.

SC
778 Tectonophysics 372, 147–166.
779
780 Cortés, M., Angelier, J., 2005. Current states of stress in the northern
781 Andes as indicated by focal mechanisms of earthquakes. Tectonophysics 403, 29–58.

U
782
783 Cortés, M., Angelier, J., Colletta, B., 2005. Paleostress evolution of the northern Andes
AN
784 (Eastern Cordillera of Colombia): Implications on plate kinematics of the South Caribbean
785 region. Tectonics 24 (27), TC1008, doi: 10.1029/2003TC001551.
786
M

787 Davis, G., Reynolds, S., Cluth, C., 2012. Structural geology of rocks and regions. Third
788 edition. John Wiley and Sons, Inc., 839pp.
789
D

790 Davy, P., Sornette, A., Sornette, D., 1990. Some consequences of a proposed fractal nature
791 of continental faulting. Nature 348, 56–58.
TE

792
793 Davy, P., Sornette, A., Sornette, D., 1992. Experimental discovery of scaling laws relating
794 fractal dimensions and the length distribution exponent of fault systems. Geophysical
795 Research Letters 19, 361–363.
EP

796
797 Davy, P., 1993. On the frequency-length distribution of the San Andreas fault system.
798 Journal of Geophysical Research 98 (12), 141–151.
C

799
800 Delvaux, D., Moeys, R., Stapel, G., Petit, C., Levi, K., Miroshnichenko, K., Ruzhich, V.,
AC

801 San’kov, V., 1997. Paleostress reconstructions and geodynamics of the Baikal region,
802 Central Asia, Part 2. Cenozoic rifting. Tectonophysics 282, 1 –38.
803
804 Delvaux, D., Sperner, B., 2003. Stress tensor inversion from fault kinematic indicators and
805 focal mechanism data: the TENSOR program. In: New Insights into Structural
806 Interpretation and Modelling (D. Nieuwland Ed.). Geological Society of London,
807 Special Publications 212, 75–100.
808
809 Del Real, C., 2013. Cartografía y Neotectónica de la Falla de Chaguacá, prolongación sur
810 de la falla de Bucaramanga. Trabajo de grado, Universidad Industrial de Santander.
811 Bucaramanga, 171pp.
812
ACCEPTED MANUSCRIPT
813 Del Real, C., Velandia, F., 2013. Cartografía geológica y evidencias de actividad reciente
814 de la Falla Chaguacá, prolongación sur de la Falla de Bucaramanga en el Macizo de
815 Floresta. XIV Congreso Colombiano de Geología, Bogotá, D.C., Memorias, 365–366.
816
817 Dengo, C., Covey, M. 1993. Structure of the Eastern Cordillera of Colombia: implications
818 for trap styles and regional tectonics. American Association of Petroleum Geologists
819 Bulletin 77 (8), 1315 ̶1337.
820
821 Dewey, J.F., 1975. Finite plate implications: Some implications for the evolution of rock

PT
822 masses at plate margins. American Journal of Science 275A, 260-284.
823
824 Diederix, H., Torres, E., Hernández, C., Botero, P., 2008. Modelo de evolución

RI
825 morfotectónica cuaternaria basado en evidencias estructurales, neotectónicas y
826 paleosimológicas de los principales sistemas de falla en la región de Bucaramanga.
827 Informe interno Ingeominas, 85pp.

SC
828
829 Diederix, H., Hernández, C., Torres, E., Osorio, J., Botero, P., 2009. Resultados
830 preliminares del primer estudio paleosimológico a lo largo de la Falla de Bucaramanga.
831 XII Congreso Colombiano de Geología, Paipa. Memorias, 18pp.

U
832
833 Doblas, M., 1998. Slickenside kinematic indicators. Tectonophysics 295 (1–2), 187–197.
AN
834
835 Doré, A., Lundin, E., Fichler, C., Olesen, O., 1997. Patterns of basement structure and
836 reactivation along the NE Atlantic margin. Journal of the Geological Society, London 154,
M

837 85-92.
838
839 Egbue, O., Kellogg, J., Aguirre, H., Torres, C., 2014. Evolution of the stress and strain
D

840 fields in the Eastern Cordillera, Colombia. Journal of South American Earth Sciences 58,
841 8-21.
TE

842
843 Fabre, A. 1983. La subsidencia de la Cuenca del Cocuy (Cordillera Oriental de Colombia)
844 durante el Cretáceo y el Terciario. Segunda parte: esquema de evolución tectónica. Revista
845 Norandina 8, 21–27.
EP

846
847 Freitas de, M., Froncolin, J., Cobbold, O., 1997. The structure of the axial zone of the
848 Cordillera Oriental, Colombia. VI Simposio Bolivariano “Exploración petrolera en las
cuencas subandinas”. Cartagena. Memorias II, 38–41.
C

849
850
AC

851 Galvis, M., Velandia, F., Villamizar, N., 2014. Cartografía morfoestructural de la Falla de
852 Bucaramanga: geometria lenticular a lo largo del valle del río Chicamocha en Santander –
853 Colombia. XVII Congreso Peruano de Geología, Memorias. Lima. Perú.
854
855 Galvis, M., 2016. Análisis estructural del segmento sur de la Falla de Bucaramanga en el
856 municipio de Onzagá, Santander. Tesis de grado, Universidad Industrial de Santander.
857 Bucaramanga, 192pp.
858
859 Giaquinta, A., Boccaletti, S., Boccaletti, M., Piccardi, P., Arecchi, F.T., 1999. Investigating
860 the Fractal Properties of Geological Fault Systems: The Main Ethiopian Rift Case.
861 Geophysical Research Letters 26, 1633–1636.
862
ACCEPTED MANUSCRIPT
863 Goltz C., 1997. Fractal and chaotic properties of earthquakes. Berlin, Springer, 178pp.
864
865 Gómez, J., Nivia, A., Montes, N., Jiménez, D., Tejada, M., Sepúlveda, M., Osorio, J.,
866 Gaona, T., Diederix, H., Uribe, H., Mora, M. 2007. Mapa y Atlas Geológico de Colombia,
867 escalas 1:2’800.000, 1:1’000.000 y 1:500.000. Ingeominas.
868
869 Gómez, J., Montes, N., Nivia. A., Diederix, H. Compiladores. 2015. Mapa Geológico de
870 Colombia, escala 1:1’000.000. Servicio Geológico Colombiano. Bogotá.
871

PT
872 González, J., Jiménez, G., 2014. Análisis estructural y características microtectónicas de un
873 segmento de la Falla de Bucaramanga en los alrededores del Corregimiento Umpalá,
874 Santander. Tesis de grado, Universidad Industrial de Santander. Bucaramanga.

RI
875
876 Guarnieri, P., Vinciguerra, S., 2004. Fractal properties of a Neogene fault system in
877 Northwestern Sicily, Italy. Fractals 12 (1), 41–48.

SC
878
879 Hastings, H.M., Sugihara, G., 1993. Fractals: A User’s Guide for the Natural Sciences.
880 Oxford Univ. Press, NY. 235pp.
881

U
882 Hentschel, H.G.E., Procaccia, I., 1983. The infinite number of generalized dimensions of
883 fractals and strange attractors, Physica DS, 435pp.
AN
884
885 Hippolyte, J. C., Bergerat, F., Gordon, M.B., Bellier, O., Espurt, N., 2012. Keys and
886 pitfalls in mesoscale fault analysis and paleostress reconstructions, the use of Angelier’s
M

887 methods. Tectonophysics 581, 144-162.


888
889 Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., Stockli, D.F.,
D

890 2010. Linking sedimentation in the northern Andes to basement configuration, Mesozoic
891 extension, and Cenozoic shortening: Evidence from detrital zircon U-Pb ages, Eastern
Cordillera, Colombia: Geological Society of America Bulletin 122 (9-10), 1423–1442.
TE

892
893
894 Ingeominas, 1997. Microzonificación sísmica del área metropolitana de Bucaramanga,
895 Santander, Colombia, Fase l. Informe preparado para la Gobernación de Santander,
EP

896 Bucaramanga, 75pp.


897
898 Irving, E.M., 1971. La evolución estructural de los Andes más septentrionales de
Colombia. Ingeominas, Boletín Geológico 19 (2), 1–90.
C

899
900
AC

901 Jiménez, G., Speranza, F., Faccena, C., Bayona, G., Mora, A., 2015. Magnetic stratigraphy
902 of the Bucaramanga alluvial fan: Evidence for a 3 mm/yr slip rate for the Bucaramanga-
903 Santa Marta Fault, Colombia. Journal of South American Earth Sciences 57, 12–22.
904
905 Kammer, A., 1996. Estructuras y deformaciones del borde oriental del Macizo de Floresta.
906 Geología Colombiana 21, 65–80.
907
908 Kammer, A. 1999. Observaciones acerca de un origen transpresivo de la Cordillera
909 Oriental. Geología Colombiana 24, 29–53.
910
ACCEPTED MANUSCRIPT
911 Kammer, A., Sánchez, J. 2006. Early Jurassic rift structures associated with the Soapaga
912 and Boyaca faults of the Eastern Cordillera, Colombia: Sedimentological inferences and
913 regional implications. Journal of South American Earth Sciences 21 (4), 412–422.
914
915 Kim, Y., Peacock, D., Sanderson, D., 2004. Fault damage zones. Journal of Structural
916 Geology 26, 503-517.
917
918 King, G., 1983. The accommodation of large strains in the upper lithosphere of the earth
919 and other solids by self-similar fault- system: the geometrical origen of b-value. Pageoph

PT
920 121, 761–815.
921
922 Koç, A., Kaymakci N., 2013. Kinematics of Sürgü fault zone (Malatya, Turkey): a remote

RI
923 sensing study. Journal of Geodynamics 65, 292-307.
924
925 Kolodyazhnyi, S.Y., 2015. The structure and evolution of the Sura–Kama strike-slip zone

SC
926 in the Cenozoic (the Volga–Ural Anteclise of the East European Platform). Geotectonics
927 49 (4), 267-288.
928
929 Mandelbrot, B.B., 1982. The Fractal Geometry of Nature. W. H. Freeman and Company.

U
930 New York, 460pp.
931
AN
932 Mandl, G., 1987. Tectonic deformation by rotating parallel faults: the “bookshelf”
933 mechanism. Tectonophysics 141, 277-316.
934
M

935 Mann, P., 2007. Global catalogue, classification and tectonic origins of restraining and
936 releasing bends on active and ancient strike-slip fault systems. In: Cunningham, W., Mann,
937 P. (Eds.), Tectonics of strike-slip restraining and releasing bends. Geological Society of
D

938 London, Special Publications 290, 13-142.


939
Marret, R., Allmendinger, R.W., 1991. Estimates of strain due to brittle faulting: sampling
TE

940
941 of fault populations. Journal of Structural Geology 13, 47–50.
942
943 McClay, K., Bonora, M., 2001. Analog models of restraining stepovers in strike-slip fault
EP

944 systems. AAPG Bulletin 85 (2), 233-260.


945
946 NASA – Alaska Satellite Facilitiy. 2015. https://vertex.daac.asf.alaska.edu/. Consultado en
junio de 2015.
C

947
948
AC

949 Naylor, M., Mandl, G., Sijpesteijn, C., 1986. Fault geometries in basement induced wrench
950 faulting under different initial stress states. Journal of Structural Geology 8 (7), 737–752.
951
952 Osorio, J., Hernández, C., Torres, E., Botero, P., 2008. Modelo geodinámico del Macizo de
953 Santander. Informe interno Ingeominas, 152pp.
954
955 París, G., Machette, M.N., Dart, R.L., Haller, K.M., 2000. Map and Database of
956 Quaternary Faults and Folds in Colombia and its Offshore Regions. A project of
957 International Lithosphere Program Task Group II-2, Major Active Faults of the World.
958 USGS. Scale 1:2’500.000. USGS Open-File Report 00-0284, 60pp.
959
ACCEPTED MANUSCRIPT
960 París, G., Romero, J., 1994. Fallas Activas de Colombia. Ingeominas, Boletín Geológico
961 34 (2-3), 3-25.
962
963 Parra, M., Mora, A., Jaramillo, C., Torres, V., Zeilinger, G., Strecker, M.R., 2010.
964 Tectonic controls on Cenozoic foreland basin development in the north-eastern Andes,
965 Colombia. Basin Research 22 (6), 874–903.
966
967 Petit, J.P., 1987. Criteria for the sense of movement on fault surfaces in brittle rocks.
968 Journal of Structural Geology 9, 597–608.

PT
969
970 Pérez-López, R., Paredes, C., Muñoz-Martín, A., 2005. Relationship between the fractal
971 dimension anisotropy of the spatial faults distribution and the paleostress fields on a

RI
972 Variscan granitic massif (Central Spain): the F-parameter. Journal of Structural Geology
973 27, 663–677.
974

SC
975 Renyi, A., 1971. Probability theory. North Holland, Amsterdam, 666pp.
976
977 Restrepo-Pace, P.A., 1995. Late Precambrian to Early Mesozoic tectonic evolution of the
978 Colombian Andes, based on new geochronological geochemical and isotopic data. PhD

U
979 Thesis. University of Arizona, USA, 195pp.
980
AN
981 Robertson, M.C., Sammis, C.G., Sahimi, M., Martin, A.J., 1995. Fractal analysis of three
982 dimensional spatial distributions of earthquakes with a percolation interpretation. Journal
983 of Geophysical Research 100, 609–620.
M

984
985 Roeder, D., Chamberlain, R., 1995. Eastern Cordillera of Colombia: Jurassic Neogene
986 crustal evolution. In: Tankard, A., Suarez, R., Welsink, H. (Eds.). Petroleum Basins of
D

987 South America. American Association of Petroleum Geologists Memoir 62, 633-645.
988
Royero, J., Vargas, R., 1999. Geología del Departamento de Santander. Mapa compilado
TE

989
990 escala 1: 300.000. Ingeominas.
991
992 Sammis, C., King, G., Biegel, R., 1987. The kinematics of gouge deformation. Pure and
EP

993 Applied Geophysics 125, 777-812.


994
995 Sanderson, D.J., Roberts, S., McGowan, A., Gumiel, P., 1991. Hercynian transpressional
tectonics at the southern margin of the Central Iberian Zone, west Spain. Journal of the
C

996
997 Geological Society 148, 893–898.
AC

998
999 Sarmiento, L., 2001. Mesozoic rifting and Cenozoic basin inversion history of the Eastern
1000 Cordillera, Colombian Andes. Inferences from tectonic models. PhD. Thesis. Vrije
1001 Universiteit Amsterdam, 296pp.
1002
1003 Sassi, W., Faure, J.L., 1996. Role of faults and layer interfaces on the spatial variations of
1004 stress regime in basins: inferences from numerical modeling. In: Cloetingh, S., Ben-
1005 Avraham, Z., Sassi, W., Horvath, F. (Eds.), Dynamics of Basin Formation and Strike-slip
1006 Tectonics. Tectonophysics 266, 101–119.
1007
ACCEPTED MANUSCRIPT
1008 Saylor, J., Horton, B., Stockli, D., Mora, A., Corredor, J., 2012a. Structural and
1009 thermochronological evidence for Paleogene basement-involved shortening in the axial
1010 Eastern Cordillera, Colombia. Journal of South American Earth Sciences 39, 202–215
1011
1012 Saylor, J.E., Stockli, D.F., Horton, B.K., Nie, J., Mora, A., 2012b. Discriminating rapid
1013 exhumation from syndepositional volcanism using detrital zircon double dating:
1014 Implications for the tectonic history of the Eastern Cordillera, Colombia. Geological
1015 Society of America Bulletin 124 (5-6), 762–79.
1016

PT
1017 Sperner, B., Zweigel, P., 2010. A plea for more caution in fault–slip analysis.
1018 Tectonophysics 482, 29–41.
1019

RI
1020 Scholz, C.H., Cowie, P., 1990. Determination of total strain from faulting using slip
1021 measurements. Nature 346, 837–839.
1022

SC
1023 Taboada, A., Rivera, L., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J.,
1024 Rivera, C., 2000. Geodynamics of the northern Andes: Subductions and Intracontinental
1025 deformation (Colombia). Tectonics 19 (5), 787–813.
1026

U
1027 Tchalenko, J. 1970. Similarities between shear zones of different magnitudes. Geological
1028 Society of America Bulletin 81, 1625–1640.
AN
1029
1030 Tesón, E., Mora, A., Silva, A., Namson, J., Teixell, A., Castellanos, J., Casallas, W.,
1031 Julivert, M., Taylor, M., Ibáñez, M., Valencia, V., 2013. Relationship of Mesozoic graben
M

1032 development, stress, shortening magnitude, and structural style in the Eastern Cordillera of
1033 the Colombian Andes. In: Nemcˇok, M., Mora, A., and Cosgrove, J.W. (Eds), Thick-Skin-
1034 Dominated Orogens: From Initial Inversion to Full Accretion. Geological Society of
D

1035 London, Special Publications 377. First published online June 11, 2013,
1036 http://dx.doi.org/10.1144/SP377.10
TE

1037
1038 Toro, J., 1990. The Termination of the Bucaramanga Fault in the Cordillera Oriental,
1039 Colombia, M.Sc. thesis. University of Arizona, 59pp.
1040
EP

1041 Toro, J., Roure, F., Bordas-Le Floch, N., Le Cornec-Lance, S., Sassi, W., 2004.
1042 Thermaland kinematic evolution of the eastern Cordillera fold-and-thrust-belt, Colombia.
1043 In: Swennen, R., Roure, F., Granath, J.W. (Eds.), Deformation, FluidFlow, and Reservoir
Appraisal in Foreland Fold and Thrust Belts. American Association of Petroleum
C

1044
1045 Geologists, Hedberg Series 1, 79–115.
AC

1046
1047 Trenkamp, R., Kellogg, J., Freymueller, J., Mora, H., 2002. Wide plate margin
1048 deformation, southern Central America and northwestern South America, CASA GPS
1049 observations. Journal of South American Earth Sciences 15, 157–171.
1050
1051 Tripathy, V., Saha, D., 2013. Plate margin paleostress variations and intracontinental
1052 deformations in the evolution of the Cuddapah basin through Proterozoic. Precambrian
1053 Research 235, 107–130.
1054
1055 Turcotte, D.L., 1986. A fractal model for crustal deformation, Tectonophysics 132, 261–
1056 269.
1057
ACCEPTED MANUSCRIPT
1058 Turcotte, D.L., 1992. Fractals and Chaos in Geology and Geophysics (Cambridge
1059 University Press, Cambridge).
1060
1061 Ujueta, G., 2003. La Falla de Santa Marta-Bucaramanga no es una sola falla; son dos fallas
1062 diferentes: la Falla de Santa Marta y la Falla de Bucaramanga. Geología Colombiana 28,
1063 133–153.
1064
1065 Ulloa, C., Guerra, A., Escovar, R., 1998. Geología de la Plancha 172 – Paz de Río, escala
1066 1:100.000. Ingeominas.

PT
1067
1068 Vargas, C., Mann, P., 2013. Tearing and breaking off of subducted slabs as the result of
1069 collision of the Panama Arc-Indenter with Northwestern South America. Bulletin of the

RI
1070 Seismological Society of America 103 (3), 2025–2046.
1071
1072 Vargas, R., Arias A., Jaramillo L., Téllez N. 1976. Geología de la Plancha 136 - Málaga,

SC
1073 escala 1:100.000. Ingeominas.
1074
1075 Vargas, R., Arias A., Jaramillo L., Téllez N. 1987. Geología de la Plancha 152 - Soatá,
1076 escala 1:100.000. Ingeominas.

U
1077
1078 Velandia, F., 2005. Interpretación de transcurrencia de las fallas Soapaga y Boyacá a partir
AN
1079 de Imágenes LANDSAT TM, Boletín de Geología 27 (1), 81–94.
1080
1081 Velandia, F., 2017. Cinemática de las fallas mayores del Macizo de Santander – énfasis en
M

1082 el modelo estructural y temporalidad al sur de la Falla de Bucaramanga. PhD Thesis.


1083 Universidad Nacional de Colombia, Bogotá, 222pp.
1084
D

1085 Veloza, G., Styron, R., Taylor, M. 2012. Open-source archive of active faults for northwest
1086 South America. GSA Today 22 (10), 4–10.
TE

1087
1088 Villamizar, N., 2014. Análisis microtectónico y morfoestructural de la Falla de
1089 Bucaramanga en el municipio de Cepitá, Santander. Tesis de grado, Universidad Industrial
1090 de Santander. Bucaramanga 118pp.
EP

1091
1092 Villar, A., Alarcón, H., Jiménez, G., Velandia, F., 2017. Zonas transversales en dominio
1093 axial de la Cordillera Oriental - bloque yacente de la Falla de Soapaga. Boletín de Geología
39 (3), 13–23.
C

1094
1095
AC

1096 Villarroel, C., Concha, A. and Macías, C., 2001. El lago Pleistoceno de Soatá (Boyacá,
1097 Colombia): Consideraciones estratigráficas, paleontológicas y paleoecológicas.
1098 Universidad Nacional de Colombia. Geología Colombiana 26, 79–93.
1099
1100 Ward, D., Goldsmith, R., Cruz, J., Jaramillo, L., Vargas, R., 1977a. Mapa geológico del
1101 cuadrángulo H-13 Pamplona, escala 1:100.000. Ingeominas.
1102
1103 Ward, D., Goldsmith, R., Jimeno, A., Cruz, J., Restrepo, H., Gómez, E., 1977b. Mapa
1104 geológico del cuadrángulo H-12 Bucaramanga, escala 1:100.000. Ingeominas.
1105
1106 Woodcock, N., Fischer, M., 1986. Strike-slip duplexes. Journal of Structural Geology 8
1107 (7), 725–735.
ACCEPTED MANUSCRIPT
1108
1109 Woodcock, N., Schubert, C., 1994. Continental strike-slip tectonics. In: Hancock, P. (Ed.),
1110 Continental deformation. Pergamon Press, New York, 251–263.
1111
1112 FIGURES AND TABLES CAPTIONS
1113
1114 Fig. 1. (a) Map of the northern part of Colombia and western Venezuela, showing the
1115 Andean cordilleras and the Maracaibo Block; GPS vectors from Trenkamp et al. (2002).
1116 (b) Geological map of the southern part of the Bucaramanga Fault in the Santander and

PT
1117 Floresta massifs. Colours modified from Gómez et al. (2015). Approximated
1118 displacements by sinistral movement along the Bucaramanga, Chaguacá and Los Micos
1119 Faults. Locations of windows of Fig. 2.

RI
1120
1121 Fig. 2. Mapping of the Bucaramanga Fault by sectors, morphologic features, and the
1122 Riedel system as indicators of sinistral strike-slip kinematics. Location of windows (a), (b),

SC
1123 (c), and (d) in Fig. 2. Digital terrain model and hillshade from NASA (2015). The Riedel
1124 faults model associated with sinistral strike-slip is taken from Woodcock and Shubert
1125 (1994) in Davis et al. (2012).

U
1126
1127 Fig. 3. (a) Sketch of the Bucaramanga Fault, defining the main trace of the strike-slip
AN
1128 structure and its southern termination as a regional restraining bend; ridges formation as a
1129 result of interaction of R with the main trace of the fault along long restraining bends; also
1130 local depressions (releasing bends) and the effect of oblique structures. (b) Idealized
1131 arrangement of a sinistral strike-slip (modified from Christie-Blick and Biddle, 1985).
M

1132
1133 Fig. 4. Map of the Bucaramanga Fault, including proposed sections and its southern
1134 termination in several branches (explanation in text). Quaternary deposits from Gómez et
D

1135 al. (2015). Digital terrain model and hillshade from NASA (2015). Stereographic
1136 projections to the lower hemisphere with the striated fault planes at each of the 30 sites;
TE

1137 principal stresses axes σ1 (circle), σ2 (triangle) and σ3 (square); σ1 orientation indicates the
1138 main compressional stress.
1139
1140 Fig. 5. Corridor of damage zone along the main trace of the Bucaramanga Fault (light
EP

1141 grey) and its southern termination in a regional restraining bend (dark grey). Stress tensors
1142 shown as arrows at each site and beach balls indicate a predominant strike-slip regime with
1143 some sites in compressional and extensional regimes, all of them with a general NW - SE
C

1144 orientation of the main compressional stress.


1145
AC

1146 Fig. 6. Mohr circles for each of the 30 deduced stress tensors. All the planes are located in
1147 the fields of neoformed or reactivated faults.
1148
1149 Fig. 7. (a) Geologic map of the southern termination of the Bucaramanga Fault and
1150 surrounding areas; colours and geological legend as in Fig. 1. (b) Detailed geologic map,
1151 modified from Vargas et al. (1976) and Velandia (2005). (c) Southern transpressive system
1152 of the Bucaramanga Fault in a restraining bend with domino pattern (lenticular) in fractal
1153 behaviour; it is shown in graded shades, where dark tones indicate the more local character
1154 of the domino structures. Numbers in circles display the names of the enunciated faults in
1155 the map legend.
1156
ACCEPTED MANUSCRIPT
1157 Fig. 8. Geological cross sections. See Fig. 7 for location. Geologic units as for the map in
1158 Fig. 7.
1159
1160 Fig. 9. North-south trending sector along the Bucaramanga fault selected for the fractal
1161 analysis. (a) 1:500,000 scale structural map across the study area, red squares represent
1162 locations of Figs. 9 (b), (c) and (d). (b) 1: 300,000 scale structural map of the northern BF.
1163 (c) 1:200,000 scale structural map of the northern BF and (d) 1: 200,000 scale structural
1164 map of the southern termination of the BF.
1165

PT
1166 Fig. 10. Variations of local fractal dimension using different box sizes for the different
1167 structural maps used (Fig. 9).
1168

RI
1169 Fig. 11. Loglog plot of numbers of boxes versus box sizes for the different structural maps
1170 scales.
1171

SC
1172
1173 Table 1. Stress tensors obtained by slickensides analysis. Parameters: n: number of faults
1174 within a subset that give the deduced tensor solution; ns: number of faults within the
1175 subset; N: number of faults measured at the site; σ1, σ2, and σ3: plunge and azimuth of

U
1176 principal stresses; α: average misfit angle between observed and modelled slip directions;
1177 F5: mean value of the optimization function; R: radio of principal stresses differences (σ2–
AN
1178 σ3)/(σ1-σ3); R': stress regime index; Shmax: direction of the maximum horizontal principal
1179 stress; QRt: stress tensor quality rank in Win-Tensor 5.8 (Delvaux and Sperner, 2003).
1180
M

1181 Table 2. Fractal dimension (mean and standard deviation) calculated across the BF system
1182 for different structural map scales (Fig. 10).
1183
D

1184
1185
TE

1186
C EP
AC
ACCEPTED MANUSCRIPT

Coordenates Stress axes


Site Age Lithology n ns N α F5 R R' Shmax QRt Stress Regime
Longitude Latitude σ1 σ2 σ3

1 -72.990 6.848 Jurassic Granite 13 42 107 20/160 65/301 14/065 7.2 2.9 0.34 1.66 158 C Pure Strike-slip

PT
2 -72.880 6.698 Paleozoic Slates 4 6 6 16/131 72/339 08/223 3 4.4 0.62 1.38 132 E Pure Strike-slip
3 -72.870 6.685 Paleozoic Quartzites 11 15 15 14/152 58/039 28/249 13.6 6.6 0.64 1.36 157 C Pure Strike-slip

RI
4 -72.810 6.614 Jurassic Quartz monzonite 9 19 19 06/349 77/106 11/258 7.5 3.9 0.26 1.74 169 D Compressional Strike-slip
5 -72.812 6.594 Jurassic Quartz monzonite 10 24 24 38/338 51/147 05/224 12.8 7.3 0.78 1.22 155 C Extensional Strike-slip

SC
6 -72.747 6.533 Cretaceous Sandstones 7 12 12 64/143 20/282 16/018 9.2 4.2 0.61 0.61 111 D Pure Extensional
7 -72.791 6.526 Jurassic Quartz monzonite 8 8 8 18/137 69/284 11/043 5.5 1.6 0.43 1.57 135 D Pure Strike-slip
8 -72.790 6.514 Jurassic Quartz monzonite 11 20 20 29/149 59/306 10/054 6.6 2 0.65 1.35 145 C Pure Strike-slip

U
9 -72.839 6.486 Jurassic Quartz monzonite 6 14 20 26/345 59/130 15/247 9.8 6.1 0.44 1.56 161 D Pure Strike-slip

AN
10 -72.799 6.491 Jurassic Quartz monzonite 12 17 17 10/289 78/144 07/020 9.9 4.3 0.30 1.70 109 C Pure Strike-slip
11 -72.755 6.491 Paleozoic Quartzites 10 15 15 28/346 62/151 06/253 8.6 5.5 0.50 1.50 164 C Pure Strike-slip
12 -72.800 6.476 Jurassic Quartz monzonite 6 6 6 21/328 68/152 02/058 6.6 1.6 0.11 1.89 148 D Compressional Strike-slip

M
13 -72.811 6.437 Jurassic Quartz monzonite 6 18 18 21/338 53/098 29/235 11.3 5.2 0.36 1.64 154 D Pure Strike-slip
14 -72.820 6.423 Jurassic Quartz monzonite 15 25 25 15/107 72/321 10/199 6.6 2.3 0.36 1.64 108 B Pure Strike-slip

D
15 -72.810 6.390 Jurassic Quartz monzonite 7 11 11 24/275 38/165 42/029 8.2 4.7 0.50 2.50 103 D Compressional Strike-slip

TE
16 -72.819 6.370 Jurassic Quartz monzonite 12 22 22 05/123 48/028 41/217 9.5 4.3 0.05 1.95 123 C Oblique Compressive
17 -72.770 6.412 Jurassic Quartz monzonite 7 10 10 32/135 58/318 0/226 6.1 2.1 0.37 1.63 135 D Pure Strike-slip
18 -72.721 6.413 Paleozoic Slates 7 15 15 35/105 52/311 13/204 7.4 3.3 0.45 1.55 110 D Pure Strike-slip
EP
19 -72.708 6.392 Cretaceous Sandstones 6 15 36 17/112 72/276 04/020 3.7 3.5 0.69 1.31 111 D Pure Strike-slip
20 -72.685 6.379 Cretaceous Sandstones 10 12 12 23/114 67/287 03/023 8.1 5.7 0.55 1.45 113 C Pure Strike-slip
C

21 -72.706 6.371 Cretaceous Sandstones 7 11 11 01/122 70/030 20/212 8.4 3.4 0.29 1.71 122 D Compressional Strike-slip
22 -72.702 6.339 Cretaceous Sandstones 22 41 41 53/109 37/300 05/206 11.9 5 0.62 0.62 115 B Pure Extensional
AC

23 -72.659 6.339 Cretaceous Sandstones 6 7 15 18/325 68/108 13/231 7.8 4.3 0.75 1.25 142 E Extensional Strike-slip
24 -72.787 6.268 Paleozoic Quartz monzonite 6 12 12 22/128 67/330 08/221 8.3 3.7 0.59 1.41 130 D Pure Strike-slip
25 -72.779 6.216 Paleozoic Quartz monzonite 18 36 36 19/292 50/178 34/035 11.8 6.7 0.15 1.85 114 B Oblique Compressive
26 -72.769 6.157 Paleozoic Conglomerates 44 110 110 38/310 43/174 23/060 9.5 4.8 0.21 1.79 135 B Oblique Extensive
27 -72.719 6.122 Cretaceous Sandstones 22 51 51 42/108 39/330 23/221 8.1 3.3 0.52 0.52 122 B Oblique Extensive
28 -72.689 6.105 Paleogene Sandstones 15 16 16 46/332 37/115 19/220 4.9 1.3 0.09 0.09 149 C Oblique Extensive
ACCEPTED MANUSCRIPT

29 -72.771 6.022 Jurassic (?) Granodiorite 10 25 25 15/128 75/316 02/218 11.5 6.7 0.51 1.49 128 D Pure Strike-slip
30 -72.754 6.010 Paleogene Sandstones 17 23 23 10/318 79/111 05/227 10.7 4.5 0.42 1.58 138 C Pure Strike-slip

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

TABLES

Manuscript The transpressive southern termination of the Bucaramanga Fault (Colombia): insights from geological mapping, stress tensor,
and fractal analysis

PT
Francisco Velandia and Mauricio Bermúdez

Table 1. Stress tensors obtained by slickensides analysis. Parameters: n: number of faults within a subset that give the deduced tensor solution; ns:

RI
number of faults within the subset; N: number of faults measured at the site; σ1, σ2, and σ3: plunge and azimuth of principal stresses; α: average misfit
angle between observed and modelled slip directions; F5: mean value of the optimization function; R: radio of principal stresses differences (σ2–

SC
σ3)/(σ1-σ3); R': stress regime index; Shmax: direction of the maximum horizontal principal stress; QRt: stress tensor quality rank in Win-Tensor 5.8
(Delvaux and Sperner, 2003).

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Coordenates Stress axes


Site Age Lithology n ns N α F5 R R' Shmax QRt Stress Regime
Longitude Latitude σ1 σ2 σ3

1 -72.990 6.848 Jurassic Granite 13 42 107 20/160 65/301 14/065 7.2 2.9 0.34 1.66 158 C Pure Strike-slip

PT
2 -72.880 6.698 Paleozoic Slates 4 6 6 16/131 72/339 08/223 3 4.4 0.62 1.38 132 E Pure Strike-slip
3 -72.870 6.685 Paleozoic Quartzites 11 15 15 14/152 58/039 28/249 13.6 6.6 0.64 1.36 157 C Pure Strike-slip
4 -72.810 6.614 Jurassic Quartz monzonite 9 19 19 06/349 77/106 11/258 7.5 3.9 0.26 1.74 169 D Compressional Strike-slip

RI
5 -72.812 6.594 Jurassic Quartz monzonite 10 24 24 38/338 51/147 05/224 12.8 7.3 0.78 1.22 155 C Extensional Strike-slip

SC
6 -72.747 6.533 Cretaceous Sandstones 7 12 12 64/143 20/282 16/018 9.2 4.2 0.61 0.61 111 D Pure Extensional
7 -72.791 6.526 Jurassic Quartz monzonite 8 8 8 18/137 69/284 11/043 5.5 1.6 0.43 1.57 135 D Pure Strike-slip
8 -72.790 6.514 Jurassic Quartz monzonite 11 20 20 29/149 59/306 10/054 6.6 2 0.65 1.35 145 C Pure Strike-slip

U
9 -72.839 6.486 Jurassic Quartz monzonite 6 14 20 26/345 59/130 15/247 9.8 6.1 0.44 1.56 161 D Pure Strike-slip

AN
10 -72.799 6.491 Jurassic Quartz monzonite 12 17 17 10/289 78/144 07/020 9.9 4.3 0.30 1.70 109 C Pure Strike-slip
11 -72.755 6.491 Paleozoic Quartzites 10 15 15 28/346 62/151 06/253 8.6 5.5 0.50 1.50 164 C Pure Strike-slip

M
12 -72.800 6.476 Jurassic Quartz monzonite 6 6 6 21/328 68/152 02/058 6.6 1.6 0.11 1.89 148 D Compressional Strike-slip
13 -72.811 6.437 Jurassic Quartz monzonite 6 18 18 21/338 53/098 29/235 11.3 5.2 0.36 1.64 154 D Pure Strike-slip

D
14 -72.820 6.423 Jurassic Quartz monzonite 15 25 25 15/107 72/321 10/199 6.6 2.3 0.36 1.64 108 B Pure Strike-slip
15 -72.810 6.390 Jurassic Quartz monzonite 7 11 11 24/275 38/165 42/029 8.2 4.7 0.50 2.50 103 D Compressional Strike-slip

TE
16 -72.819 6.370 Jurassic Quartz monzonite 12 22 22 05/123 48/028 41/217 9.5 4.3 0.05 1.95 123 C Oblique Compressive
17 -72.770 6.412 Jurassic Quartz monzonite 7 10 10 32/135 58/318 0/226 6.1 2.1 0.37 1.63 135 D Pure Strike-slip
EP
18 -72.721 6.413 Paleozoic Slates 7 15 15 35/105 52/311 13/204 7.4 3.3 0.45 1.55 110 D Pure Strike-slip
19 -72.708 6.392 Cretaceous Sandstones 6 15 36 17/112 72/276 04/020 3.7 3.5 0.69 1.31 111 D Pure Strike-slip
C

20 -72.685 6.379 Cretaceous Sandstones 10 12 12 23/114 67/287 03/023 8.1 5.7 0.55 1.45 113 C Pure Strike-slip
AC

21 -72.706 6.371 Cretaceous Sandstones 7 11 11 01/122 70/030 20/212 8.4 3.4 0.29 1.71 122 D Compressional Strike-slip
22 -72.702 6.339 Cretaceous Sandstones 22 41 41 53/109 37/300 05/206 11.9 5 0.62 0.62 115 B Pure Extensional
23 -72.659 6.339 Cretaceous Sandstones 6 7 15 18/325 68/108 13/231 7.8 4.3 0.75 1.25 142 E Extensional Strike-slip
24 -72.787 6.268 Paleozoic Quartz monzonite 6 12 12 22/128 67/330 08/221 8.3 3.7 0.59 1.41 130 D Pure Strike-slip
25 -72.779 6.216 Paleozoic Quartz monzonite 18 36 36 19/292 50/178 34/035 11.8 6.7 0.15 1.85 114 B Oblique Compressive
26 -72.769 6.157 Paleozoic Conglomerates 44 110 110 38/310 43/174 23/060 9.5 4.8 0.21 1.79 135 B Oblique Extensive
27 -72.719 6.122 Cretaceous Sandstones 22 51 51 42/108 39/330 23/221 8.1 3.3 0.52 0.52 122 B Oblique Extensive
28 -72.689 6.105 Paleogene Sandstones 15 16 16 46/332 37/115 19/220 4.9 1.3 0.09 0.09 149 C Oblique Extensive
29 -72.771 6.022 Jurassic (?) Mylonite 10 25 25 15/128 75/316 02/218 11.5 6.7 0.51 1.49 128 D Pure Strike-slip
30 -72.754 6.010 Paleogene Sandstones 17 23 23 10/318 79/111 05/227 10.7 4.5 0.42 1.58 138 C Pure Strike-slip
ACCEPTED MANUSCRIPT

Table 2. Fractal dimension (mean and standard deviation) calculated across the BF system for different structural map scales (Fig. 10).

PT
Structural map scale Fractal dimension (Df) mean Standard deviation (s)
Map 1:500,000 1.4159 0.071958

RI
Map 1:300,000 1.4079 0.092542
Map 1:200,000 (North) 1.3655 0.083294

SC
Map 1:200,000 (South) 1.3931 0.12603

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
FIGURES

Manuscript The transpressive southern termination of the Bucaramanga Fault


(Colombia): insights from geological mapping, stress tensors, and fractal analysis
Francisco Velandia and Mauricio Bermúdez

Fig. 1. (a) Map of the northern part of Colombia and western Venezuela, showing the Andean cordilleras and
the Maracaibo Block; GPS vectors from Trenkamp et al. (2002). (b) Geological map of the southern part of
the Bucaramanga Fault in the Santander and Floresta massifs. Colours modified from Gómez et al. (2015).
Approximated displacements by sinistral movement along the Bucaramanga, Chaguacá and Los Micos Faults.

PT
Locations of windows of Fig. 2.

RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 2. Mapping of the Bucaramanga Fault by sectors, morphologic features, and the Riedel system as
indicators of sinistral strike-slip kinematics. Location of windows (a), (b), (c), and (d) in Fig. 2. Digital terrain
model and hillshade from NASA (2015). The Riedel faults model associated with sinistral strike-slip is taken
from Woodcock and Shubert (1994) in Davis et al. (2012).

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 3. (a) Sketch of the Bucaramanga Fault, defining the main trace of the strike-slip structure and its southern
termination as a regional restraining bend; ridges formation as a result of interaction of R with the main trace
of the fault along long restraining bends; also local depressions (releasing bends) and the effect of oblique
structures. (b) Idealized arrangement of a sinistral strike-slip (modified from Christie-Blick and Biddle, 1985).

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 4. Map of the Bucaramanga Fault, including proposed sections and its southern termination in several
branches (explanation in text). Quaternary deposits from Gómez et al. (2015). Digital terrain model and
hillshade from NASA (2015). Stereographic projections to the lower hemisphere with the striated fault planes
at each of the 30 sites; principal stresses axes σ1 (circle), σ2 (triangle) and σ3 (square); σ1 orientation indicates
the main compressional stress.

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 5. Corridor of damage zone along the main trace of the Bucaramanga Fault (light grey) and its southern
termination in a regional restraining bend (dark grey). Stress tensors shown as arrows at each site and beach
balls indicate a predominant strike-slip regime with some sites in compressional and extensional regimes, all
of them with a general NW - SE orientation of the main compressional stress.

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 6. Mohr circles for each of the 30 deduced stress tensors. All the planes are located in the fields of
neoformed or reactivated faults.

Part 1

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Part 2

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 7. (a) Geologic map of the southern termination of the Bucaramanga Fault and surrounding areas; colours
and geological legend as in Fig. 1. (b) Detailed geologic map, modified from Vargas et al. (1976) and Velandia
(2005). (c) Southern transpressive system of the Bucaramanga Fault in a restraining bend with domino pattern
(lenticular) in fractal behaviour; it is shown in graded shades, where dark tones indicate the more local character
of the domino structures. Numbers in circles display the names of the enunciated faults in the map legend.

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 8. Geological cross sections. See Fig. 7 for location. Geologic units as for the map in Fig. 7.

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 9. North-south trending sector along the Bucaramanga fault selected for the fractal analysis. (a) 1:500,000
scale structural map across the study area, red squares represent locations of Figs. 9 (b), (c) and (d). (b) 1:
300,000 scale structural map of the northern BF. (c) 1:200,000 scale structural map of the northern BF and (d)
1: 200,000 scale structural map of the southern termination of the BF.

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 10. Variations of local fractal dimension using different box sizes for the different structural maps used
(Fig. 9).

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Fig. 11. Loglog plot of numbers of boxes versus box sizes for the different structural maps scales.

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

HIGHLIGHTS

Manuscript The transpressive southern termination of the Bucaramanga Fault


(Colombia): insights from geological mapping, stress tensors, and fractal analysis

Francisco Velandia and Mauricio Bermúdez

PT
• The southern termination of the fault shows a transpressive system in domino
style

RI
• Tensor analysis indicate a NW-SE orientation for the maximum horizontal stress

• Morphostructures and stress tensor analysis confirm sinistral kinematics

SC
• The positive flower structure of the fault termination presents fractal behavior

U
• Prior its termination the fault develops a wide damage zone of about 8 km width
AN
M
D
TE
C EP
AC

You might also like