Fault Characterization and Heat-Transfer Modeling To The Northwest of Nevado Del Ruiz Volcano

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 40

Accepted Manuscript

Fault characterization and heat-transfer modeling to the Northwest of Nevado del


Ruiz Volcano

David Moreno Rendon, Jacqueline Lopez-Sanchez, Daniela Blessent, Jasmin


Raymond

PII: S0895-9811(18)30068-3
DOI: 10.1016/j.jsames.2018.08.008
Reference: SAMES 1980

To appear in: Journal of South American Earth Sciences

Received Date: 17 February 2018


Revised Date: 7 August 2018
Accepted Date: 10 August 2018

Please cite this article as: Rendon, D.M., Lopez-Sanchez, J., Blessent, D., Raymond, J., Fault
characterization and heat-transfer modeling to the Northwest of Nevado del Ruiz Volcano, Journal of
South American Earth Sciences (2018), doi: 10.1016/j.jsames.2018.08.008.

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 Fault characterization and heat-transfer modeling to the Northwest of Nevado del Ruiz
2 Volcano

3 David Moreno Rendona, Jacqueline Lopez-Sancheza, Daniela Blessenta, Jasmin Raymondb

4
a

PT
5 Universidad de Medellín, Programa de Ingeniería Ambiental, Medellín, Colombia
b
6 Institut national de la recherche scientifique, Centre Eau Terre Environnement, Québec, Qc,

RI
7 Canada

SC
9

10

U
11
AN
12

13
M

14
D
TE
C EP
AC

1
ACCEPTED MANUSCRIPT

15 Abstract

16 The Nevado del Ruiz Volcano is an area of great interest for future geothermal exploitation in

17 Colombia facing exploration challenges as hydrothermal fluids originate deep into the fractured

18 basement. Fieldwork conducted on the Northwest of this volcano confirmed the existence of fault

PT
19 zones and enabled collection of thirty rock samples from outcrops. Permeability of these samples

20 was then measured in the laboratory with a steady-state gas permeameter, taking into account the

RI
21 Klinkenberg correction. The Porchet method allowed to obtain an estimation of in situ hydraulic

SC
22 conductivity in fault zones directly in the field. Thermal conductivity and heat capacity of the

23 rock units were estimated from previous studies. 2D heat transfer and groundwater flow in porous

U
24 fractured medium were simulated in the study area, using the free modeling package

25 OpenGeoSys with data collected during fieldwork.


AN
26 Three modeling scenarios were considered, investigating the influence of fault dip in groundwater

27 flow and heat transfer processes. It was possible to reproduce the temperature measured at the
M

28 Hotel Termales hot spring, where the Samaná Sur fault appears to behave as a pathway for
D

29 geothermal fluids. It was additionally found that fault dip has an influence on the simulated
TE

30 temperatures. When the dip is in favor of the fluid flow (<90° facing fluid flow), the simulated

31 surface temperature increases; on the other hand, if the dip of fault is opposite to the fluid flow (>
EP

32 90° facing fluid flow), the simulated surface temperature decreases.

33
C

34 Keywords:
AC

35 Nevado del Ruiz Colombia, fault zones, OpenGeoSys, heat transfer, groundwater flow,
36 geothermal energy

37

38

2
ACCEPTED MANUSCRIPT

39 1. Introduction

40 Faults and associated fractures are important for the understanding of geothermal systems due to

41 their influence on stratigraphic structure and groundwater flow (Corbel et al., 2012). The fault

42 location, orientation, geometry and magnitude of permeability contrast between the fault core and

PT
43 the damage zone define the behavior of faults as a pathway, barrier, or both, for geothermal fluids

44 (Caine et al., 1996; Li et al., 2013). Usually, a well-developed fault core cataclasites and an

RI
45 absent to poorly developed damage zone will induce a localized barrier to the fluid flow, while a

SC
46 well-developed and associated fractured networks will induce a distributed conduit (Caine et al.,

47 1996). If faults behave as conduits and increase the permeability (Ferrill et al., 2004), they can be

U
48 a target for geothermal drilling, but if they behave as barriers and decrease the permeability

49 (Gibson, 1998), they can cause compartmentalization of reservoirs, which can generate
AN
50 difficulties for access and exploitation of the geothermal resources (Lovelessa et al., 2014).

51 Holzbecher et al. (2010) and Kalinina et al. (2013) further discussed the importance of trends and
M

52 dips of faults and fractures on fluid circulation. Numerical modeling has been used as a tool to
D

53 better understand fractured geothermal reservoirs (Holzbecher et al., 2010; Corbel et al., 2012;
TE

54 Hao et al., 2012; Bakhsh et al., 2016). The models to represent fractures can be grouped into two

55 categories: Fracture Continuum Models (FCM) and Discrete Fracture Models (DFM). The FCM
EP

56 is based on the concept that the fractured rock can be treated as a continuum when the volume of

57 the rock is large enough for this approximation. On the other hand, the DFM represents explicitly
C

58 the fractures in the model (Jaffre et al., 2011) and makes it possible to analyze the physical
AC

59 processes in each fracture (Li et al., 2013). Prediction and modeling of coupled fluid flow and

60 heat transfer processes in naturally and fractured rock systems represent a critical component for

61 energy recovery analysis (Hao et al., 2012). The influence of faults should be taken into account

62 when exploring geothermal reservoirs, like at the Nevado del Ruiz (NDR) geothermal project in

63 Colombia. Many faults have been identified in this area and are indicated on geological maps

64 (Mosquera et al., 1998a, 1998b; Martínez et al., 2014; Gómez et al., 2015).

3
ACCEPTED MANUSCRIPT

65 It is therefore of great interest to understand the role played by those faults in heat transfer

66 processes for the characterization and evaluation of geothermal resources (Pérez et al., 2012;

67 Sanchez-Alfaro et al., 2015; Aravena et al., 2016; Benavente et al., 2016).

68 The NDR volcano is the best known geothermal area of Colombia, where the interest of several

PT
69 entities converge (Alfaro, 2015). Colombia is, however, lagging behind other countries with

70 similar volcanic environments, when comparing geothermal power development in Latin

RI
71 America (Bertani, 2016). Actually, there are no models of groundwater flow and heat transfer for

SC
72 the geothermal reservoir of the NDR showing the influence of fault zones on the simulated

73 temperature. Faults are expected to control hydrothermal fluid circulation in the NDR geothermal

U
74 reservoir (González-García et al., 2015) and they should be taken into account for a detailed

75 estimation of the geothermal potential (Vélez et al., 2018). The objective of the work described in
AN
76 this paper is to better understand the impact of faults on fluid circulation from the basement

77 underlying the NDR volcanic complex to the surface, where hot springs manifestations are
M

78 observed. Concepts developed here could be used to better understand other hydrothermal
D

79 systems found in similar geological setting of the Andes. A 2D numerical groundwater flow and
TE

80 heat transfer model of an area crossed by several faults located to the Northwest of the NDR was

81 developed for that purpose. The numerical model has been built from 1) petrographic thin section
EP

82 descriptions, 2) permeability and density laboratory measurements on rock samples collected at

83 outcrops, 3) field measurement of hydraulic conductivity with the Porchet method, strike and dip
C

84 measurements of faults, 4) inferred geological cross-section, and 5) numerical modeling of porous


AC

85 fractured medium with the OpenGeoSys software (Böttcher et al., 2016), considering DFM for

86 fault representation. The work presented here can be seen as the logical continuation of the first

87 groundwater flow and heat transfer numerical model of the NDR area presented by Vélez et al.

88 (2018) with the introduction of new concepts based on the field and numerical modeling analysis

89 of fault zones.

90

4
ACCEPTED MANUSCRIPT

91 2. Geological setting of the Nevado del Ruiz Volcano

92 Trenkamp et al. (2002) mentioned that volcanism in the Colombian Central Andes is defined by a

93 complex tectonic framework, since the country is located at the intersection of South America,

94 Nazca, and Caribbean tectonic plates (Figure 1), together with the Coiba microplate. The

PT
95 subduction of the Nazca beneath the South American Plate is the governing mechanism triggering

96 volcanic activity due to the rapid convergence of 58 mm/year occurring at the Colombia-Ecuador

RI
97 trench (Trenkamp et al., 2002).

SC
98 The NDR (5321 m.a.s.l) is an ice-capped stratovolcano enclosing the Los Nevados National

99 Natural Park (NNP) located at the northern end of the Northern Volcanic Zone in the central

U
100 Andes mountain range. Its geographical position is 4.89°N and 75.32°W, 140 km NW of the
AN
101 capital Bogotá and 28 km SE of Manizales city. Along with Puracé and Galeras volcanoes, it is

102 one of the most active volcanoes in Colombia (Stix et al., 2003). The eruptive history of the NDR
M

103 volcano starts 1.8 million years ago in the Pliocene and it is subdivided in 3 stages: Ancestral

104 Ruiz (1.8-0.8 m.y), Ancient Ruiz (0.8-0.1 m.y), and Recent Ruiz (0.1 m.y to present), as
D

105 described by Thouret (1990).


TE
C EP
AC

106

5
ACCEPTED MANUSCRIPT

107 Figure 1. Tectonic framework close to the Nevado del Ruiz, which is represented by the red

108 triangle (Trenkamp et al., 2002; Martínez et al., 2014)

109 The study area is located to the Northwest of the NDR (Figure 2) and its geology is described by

110 González (2001) and resumed here with its main characteristics. The Quebradagrande Complex is

PT
111 located to the West of the study area and is characterized by Northwest intercalation between

RI
112 Volcanic Complex (Kvc) and Sedimentary Complex (Ksc), both with dynamic metamorphism,

113 where the contact zone is represented by a weakness area that affects the Sedimentary Complex

SC
114 with more intensity due its hardness. The Volcanic Complex is composed of basalts, pyroclastic

115 flows and diabasic dykes, while the Sedimentary Complex is composed of black shales,

U
116 sandstones, conglomerates, and limestones. The Andesite unit (NgQa) is located to the East of the
AN
117 study area and has a composition ranging from andesitic to dacitic, and basaltic at fewer

118 locations. The Cajamarca Complex is located in the middle and to the East of the study area and it
M

119 is a set of metamorphic rocks, such as phyllites and schist composed principally by sericite,

120 quartz, graphite, chlorite, muscovite, biotite, and other minerals. This Complex is the result of
D

121 modified mineralogy due to the diverse regional metamorphism increasing temperature and
TE

122 pressure conditions with variable intensity (González, 2001). A set of compressive stress is

123 caused by oblique convergence between Nazca and South America plates that produce stress
EP

124 tensors with regional variability. These stresses cause differential reactivation of fault systems

125 parallel to the Andes mountain range, exhibiting a transpressive regime (Toro and Osorio, 2005).
C

126 The five fault systems in the study area are the Villamaria-Termales, Nereidas-Rio Claro,
AC

127 Romeral, Palestina, Pico Terrible-Samaná Sur-Olletas fault systems (Figure 2).

128 The Villamaria-Termales fault system was initially described and defined by Thouret (1988) as

129 sinistral strike-slip faults. The Villamaria-Termales fault system is a series of normal faults that

130 trends WNW-ESE and are located in the North of the NDR (Mejía et al., 2012; González-García

131 and Jessell, 2016).

6
ACCEPTED MANUSCRIPT

132 The Nereidas and Rio Claro faults are defined as normal faults with left lateral component

133 dipping to NE with a NW-SE trend (Ceballos, 2017). They are characterized by the presence of

134 hot springs and by their structural control of the river channels, from which their names are

135 derived (CHEC et al., 1983).

PT
136 The Romeral fault system is a group of parallel faults located further West of the NDR with N-S

137 and a NNE-SSW trend. Silvia Pijao, Manizales, and San Jeronimo faults are some of the most

RI
138 important faults of this system. Silvia Pijao and Manizales faults are defined as reverse faults with

SC
139 a dextral slip component dipping East (Camargo and Mojica 2004; Mejía et al., 2012; González-

140 García and Jessell, 2016). The San Jeronimo fault divides the Quebradagrande Complex from the

U
141 Cajamarca Complex. It is a reverse fault with left lateral displacement (Martínez et al., 2014) and

142 it represents the farther east limit of the Romeral fault system.
AN
143 The Palestina fault system passes through the Ruiz-Tolima volcanic complex holding a high

144 structural control on it with a NNE-SSW and ENE-WSW trend. It shows alignment with focal
M

145 volcanic points of NDR, El Cisne and Santa Isabel, and it is characterized by right lateral strike-
D

146 slip faults.


TE

147 The Pico Terrible, Samaná Sur and Olletas faults show the same tendencies according to

148 Martínez et al. (2014). Samaná Sur is defined as a reverse fault with a right lateral component,
EP

149 trending NE-SW (González, 1980) and dipping East. Martínez et al. (2014) mentioned that the

150 Olleta-Nereidas fault is characterized by a strike-slip regime and that Samaná Sur fault is
C

151 characterized by three separate segments where hot springs and other hydrothermal activities are
AC

152 observed.

153 Several recharge zones and hot springs close to the NDR are caused by the interaction of different

154 fault systems, mainly in dilatational tectonic syntax zone represented by the intersection between

155 two or more strike-slip faults with opposite movements, and along faults running NW-SE and

156 their interaction with faults running NS and NE-SW (Mejía et al., 2012; Forero, 2012; Ceballos,

157 2017).

7
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
158
M

159 Figure 2. Structural map of the study area, showing hot springs and faults. Modified from the

160 sheets # 206 and # 225 of the INGEOMINAS geological map (Mosquera et al., 1998a, 1998b)
D

161 and from the geology and stratigraphy of the NDR volcanic complex map (Martínez et al., 2014).
TE

162

163 3. Previous studies


EP

164 The first geothermal study on the NDR site was conducted at the end of the sixties by the Italian

165 company ENEL (Ente Nazionale per la Energia Elettrica) in collaboration with the CHEC
C

166 (Central Hidroeléctrica de Caldas) and described litho-stratigraphic features, volcanology,


AC

167 structural events and hydrogeology of the NDR complex. The presence of a thick caprock,

168 formed by the upper part of the metamorphic complex and separating the two major circulation-

169 systems to the West, North-West and North of the Ruiz edifice was identified (Arango et al.,

170 1970).

8
ACCEPTED MANUSCRIPT

171 In 1983, the CHEC performed prefeasibility studies for geothermal development in the NDR

172 Volcanic Complex, providing a detailed interpretation of the origin and evolution of the

173 hydrothermal system associated with the volcano. This characterization included the

174 determination of pH, conductivity, and temperature and other properties of and components of the

PT
175 hydrothermal system (CHEC et al., 1983).

176 The only geothermal exploratory well drilled in Colombia (Nereidas 1) is located on the western

RI
177 flank of Nevado del Ruiz volcano and was conducted by GEOENERGIA ANDINA S.A. (GESA)

SC
178 in 1997. The well was drilled to 1468 m depth, through seven lithologic units with hydrothermal

179 alteration. The measured bottom-hole temperature was about 200°C; however the study suggested

U
180 a geothermal reservoir temperature higher than 200°C taking into account the presence of epidote

181 in the basement rocks as an indicator of temperatures above 250°C. The study also suggested that
AN
182 contact metamorphism veins of adularia and albite in schists and minerals like pyrite indicate

183 secondary permeability in the basement rocks and consequently in the reservoir (Monsalve et al.,
M

184 1998).
D

185 Londoño and Sudo (2002) performed a seismic tomographic study at the NDR. The study
TE

186 suggested the presence of two magmatic chambers located to a depth of 5 and 10 km,

187 respectively, which might be intercepted by Villamaria-Termales and Palestina fault systems.
EP

188 Recently, Rojas (2012) presented the temperature profiles measured in three 300 m deep wells

189 that were drilled in 2011 with the objective to measure the geothermal gradient. Forero (2012)
C

190 provided a characterization of hydrothermal alteration to the NW of the volcano that led to a
AC

191 simple conceptual model of the geothermal reservoir, which highlighted the importance of fault

192 systems as pathways for hot fluids. Mejía et al. (2012) performed structural geology studies

193 around the NDR, based on digital elevation models and analysis of morphological and tectonic

194 features. The study proposed a structural scheme that explains the pattern of location of hot

195 springs and fluid released in the western part of the NDR. Almaguer (2013) presented the results

196 from a magnetotelluric study conducted to the North of the volcano, where high electrical

9
ACCEPTED MANUSCRIPT

197 conductivity areas confined by sections of moderated resistivity were interpreted as a possible

198 reservoir and caprock, respectively. Martínez et al. (2014) published a geological map of Nevado

199 del Ruiz Volcanic Complex with an accurate description of stratigraphy and fault systems.

200 González-García and Jessell (2016) published a first 3D geological model for the whole Ruiz-

PT
201 Tolima volcanic massif, using the Monte Carlo method to characterize geological uncertainty.

202 Their geological model represents the probability of occurrence of geological units, suggesting

RI
203 where future exploratory work should be conducted.

SC
204

205 4. Material and methods

U
206 4.1 Study area, rock sampling and Porchet method

207 The Colombian National Seismic Network (RSNC) and the Colombian Geological Survey (SGC)
AN
208 lead the seismological investigations in Colombia and are in charge of giving warning during the

209 occurrence of a seismic event in the national territory. There are 53 satellite seismological
M

210 stations (long and short term) and 111 accelerograph stations, the NDR being one of the most
D

211 monitored volcanoes in the world since an eruption that caused up to 25,000 casualties occurred
TE

212 in 1985 (El Espectador, 2014). Using the available data from RSNC and SGC, and according

213 Spence et al. (1989) that classify shallow earthquakes between 0 and 70 km deep and
EP

214 intermediate/deep earthquakes > 70 km deep, a seismic analysis distinguishing shallow-focus

215 from deep-focus earthquakes was conducted. Earthquakes alignments were inferred taking into
C

216 account the alignments of the shallow focal depths seismic data as evidence of fault activity in the
AC

217 study area (Spence et al., 1989). A geological cross section was drawn through the active faults

218 identified in this seismic analysis.

219 Fieldwork to the nNorthwest of the NDR was conducted in June 2016 near the town of

220 Villamaria. The geological cross-section AA’ intersects the Quebradagrande Complex (Ksc and

221 Kvc) and the Cajamarca Complex (Pes), which is the lithology of most interest, because it is

222 considered the potential geothermal reservoir despite zones with primary porosity have not been

10
ACCEPTED MANUSCRIPT

223 identified (Almaguer, 2013). The Cajamarca Complex (Pes) is deep enough to host fluids of

224 sufficient temperature for geothermal power generation (Vélez et al., 2018). The cross-section has

225 an approximate length of 21 km and starts from Silvia Pijao fault to the West (A) and goes to the

226 Termales del Ruiz Hotel to the East (A’), which is the closest part to the NDR volcano (Figure 3).

PT
RI
U SC
AN
M
D
TE
C EP
AC

227
228 Figure 3. Geological map of the study area showing the location of geological cross-section AA’,

229 hot springs, and fieldwork sampling points. Modified from the sheets #206 and #225 of the

230 INGEOMINAS geological map (Mosquera et al., 1998a, 1998b) and from the geology and

231 stratigraphy of the NDR volcanic Complex map (Martínez et al., 2014).

11
ACCEPTED MANUSCRIPT

232 The objective of the fieldwork was to identify and characterize 9 different fault zones in the study

233 area. 31 rock samples belonging to four different lithologies were collected including three

234 replicas for each sampling point (Table 1), both from fractured and no fractured outcrops, with

235 the intention to evaluate in the laboratory hydraulic properties for numerical modeling.

PT
236

237 Table 1. Sampling point location (WGS84 coordinate system)

RI
Code
P: point X coordinate Y coordinate Geological unit
M: sample

SC
Quebradagrande
P1M1 -75.558698 4.979891 Sedimentary
Complex (Ksc)
Quebradagrande

U
P1M2 -75.558698 4.979891 Volcanic
Complex (Kvc)
AN
Quebradagrande
P2M1 -75.551064 4.986907 Volcanic
Complex (Kvc)
M

Quebradagrande
P3M1 -75.539246 4.98052 Sedimentary
Complex (Ksc)
D

Quebradagrande
P3M2 -75.539246 4.98052 Volcanic
TE

Complex (Kvc)
Quebradagrande
P5M1 -75.503881 4.966493 Volcanic
Complex (Kvc)
EP

Quebradagrande
P6M1 -75.491988 4.985406 Volcanic
Complex (Kvc)
C

Cajamarca
P7M1 -75.487735 4.975936
Complex (Pes)
AC

Cajamarca
P8M1 -75.456529 4.97659
Complex (Pes)
Cajamarca
P9M1 -75.453491 4.975615
Complex (Pes)
Andesite unit
P11M1 -75.379769 4.970774
(NgQa)

12
ACCEPTED MANUSCRIPT

238 The Porchet method is described by Van der Molen et al. (2007) and applied in several works

239 such as Colombani et al. (1972) and Puga-Lagos (2012) as a field method for measuring

240 hydraulic conductivity, mainly in unconsolidated sediments. However, taking into account the

241 difficulty to evaluate hydraulic conductivity for fault zones and the importance of this parameter

PT
242 in the characterization of the fractured geothermal reservoir, the Porchet method was used to get

243 an estimate of the hydraulic conductivity of fault zones at the field scale.

RI
244 A shallow hole with known depth and diameter was drilled in outcrops with a horizontal surface

SC
245 according to the Porchet method. Then, water was poured into the hole and the rate at which the

246 water level decreased was measured (Figure 4). Finally, Eq.(1), recommended by Oosterbaan and

U
247 Nijland (1994), was used to calculate the hydraulic conductivity (Van der Molen et al., 2007) of

248 the fractured formations.


AN
 r
r  h1 + 2 
249 K= ln   h1 > h2 ; t2 > t1 (1)
2 ( t2 − t1 )  h + r 
M

 2
2

250 where h1 [L] and h2 [L] are the initial and final water level in the hole, respectively, t1 and t2 [T]
D

251 are the initial and final times, r is the hole radius [L] and K [LT-1] is the hydraulic conductivity. In
TE

252 this case, the Porchet method was used to interpret the infiltration tests conducted in the outcrops,

253 along fault zones, where rocks were fractured enough to dig holes.
C EP
AC

254

255 Figure 4. Infiltration test analyzed with Porchet method to determine hydraulic conductivity of

256 fractured rocks.

13
ACCEPTED MANUSCRIPT

257 4.2 Thin sections, hydraulic and physical properties

258 Four rock samples of Kvc, two of Ksc, five of Pes and one of NgQa were sent to GMASLAB

259 S.A.S laboratory in Bogotá, Colombia, to evaluate permeability and grain density, and to conduct

260 thin section analysis. The Klinkenberg correction was applied to evaluate the liquid permeability

PT
261 in the porous media, which can differ from the gas permeability due to gas slippage effect. The

262 Klinkenberg correction can be achieved by the injection of gas with different pressures that will

RI
263 flow through the porous medium in order to evaluate slippage effect. The procedures and

SC
264 measurements for the evaluation of permeability and grain densities were made according to API-

265 RP40/98 (American Petroleum Institute, 1998). The hydraulic conductivity was calculated

U
266 considering the water kinematic viscosity at the temperature of the river, stream or hot spring
AN
267 close to the sampling zone. The temperature from the closest river (21°C) was selected for the

268 Kvc and Ksc samples; the temperature of the Hacienda la Quinta hot springs (100°C) was
M

269 considered for the Pes sample 2014MI21 collected close to it; the temperature of the closest river

270 (15°C) was used to calculate hydraulic conductivity of the other Pes samples; finally, the
D

271 temperature from the Hacienda Termales hot spring was selected (63°C) for the NgQa sample.
TE

272 Considering the surface water manifestation as representative of the average temperature of the

273 hydrothermal fluid in the liquid phase was assumed appropriate for the modeling scale and
EP

274 objectives of this study.

275 The fluid flow through fractures in the DFM was assumed to be governed by Poiseuille’s law,
C

276 also known as the cubic law (Witherspoon et al., 1980), which is an analytical solution for
AC

277 laminar flow between smooth parallel plates (Li et al., 2013). Taking into account Eq. (2),

278 fracture aperture was estimated:

γ ( 2b )
2

279 K fr = (2)
µ 12

280 where Kfr [LT-1] is the fracture hydraulic conductivity, γ [FL-3] is the water specific weight , µ

281 [FL-2T] is the water dynamic viscosity, 2b [L] is the fracture aperture. The hydraulic conductivity

14
ACCEPTED MANUSCRIPT

282 was measured in the field with the Porchet method in the identified fault zones. Since it was

283 difficult to identify the real aperture of the fault zones, it was estimated using Eq. (2) assuming

284 that the discrete fracture in the numerical model approximates the behavior of the real fault zone,

285 which was considered suitable for the modeling scale and the objectives of this work.

PT
286

287 4.3 Numerical modeling

RI
288 The software OpenGeoSys (OGS), a scientific open-source initiative for numerical simulation of

SC
289 thermo-hydro-mechanical/chemical processes in porous and fractured media (Böttcher et al.,

290 2016), was used to build a 2D numerical model based on the geological cross-section (AA’). The

U
291 modeling objective was to estimate underground temperature and identify the influence of faults

292 on the simulated temperature in the geothermal reservoir. The DFM approach was selected to
AN
293 represent the faults in the numerical model.

294 The GMSH mesh generator (Geuzaine and Remacle, 2009) and the Paraview software were used
M

295 to build the mesh and to visualize the numerical results, respectively. Fully-saturated steady-state
D

296 groundwater flow and transient-state heat transfer were the physical processes considered.
TE

297 Governing equations and specifications for modeling of geothermal processes are presented in

298 detail by Böttcher et al. (2016) and therefore they are not repeated here.
EP

299

300 5 Results
C

301 5.1 Seismic analysis


AC

302 More than a half of the earthquakes in the study area have focal depths less than 5 km, which are

303 classified as shallow focal depths (< 70 km) and are caused by active faults, while there are few

304 deep-focus earthquakes (> 70 km) that occur only in association with subduction zones along

305 convergent plate boundaries (Spence et al., 1989), as shown in Figure 5A. Most of the earthquake

306 alignments coincide with fault location and the identification of different alignment intersections

307 allowed to recognize recent fractured zones and active faults. According to the last analysis, nine

15
ACCEPTED MANUSCRIPT

308 important faults (two of them inferred) were selected for field characterization, while only seven

309 were included in the groundwater flow and heat transport model (Table 2), as the two inferred

310 faults were neglected given their existence and location and existence are uncertain.

PT
RI
U SC
AN
M
D
TE
C EP
AC

311

312 Figure 5. A) Shallow and intermediate/deep earthquakes. B) Location of seismic stations to the

313 Northwest of NDR. Maps modified from the sheets #206 and #225 of the INGEOMINAS

16
ACCEPTED MANUSCRIPT

314 geological map (Mosquera et al., 1998a, 1998b), from the geology and stratigraphy of the NDR

315 volcanic Complex map (Martínez et al., 2014), and from available seismological data from RSNC

316 and SGC.

317 5.2 Field measurements

PT
318 Several hot springs reported by the Colombian Geological Survey were dry during fieldwork due

319 to limited precipitations. However, the presence of hydrothermal alteration minerals like epidote

RI
320 close to the fault zones and far from the NDR is evidence that those locations hosted hot springs

SC
321 in the past. This was observed with sample P3M2 (Kvc) collected to the West of the study area.

322 The radius and depth of the holes dug for infiltration tests with the Porchet method were not

U
323 constant in all sampling zones due to the difficulty to dig the rock. It was not possible to reach the

324 Samaná Sur and Pico Terrible faults to conduct the tests because of lack of access routes. The
AN
325 value for these faults were estimated assuming the higher aperture and hydraulic conductivity due

326 to the presence of several hot springs, which were not observed close to other fault zones. Dip of
M

327 faults was measured and compared with the information from the geologic map of NDR Volcanic
D

328 Complex (Martínez et al., 2014).


TE

329 Table 2. Properties of faults used for numerical simulations


Hydraulic
Aperture Transmissivity
Dip (°) Conductivity
-1 (m) (m2 s-1)
(m s )
EP

Fault_0 56NE 2.03 x 10-4 1.69 x 10-5 3.43 x 10-9

Fault_1
C

54NE 6.96 x 10-5 9.88 x 10-6 6.88 x 10-10


(Silvia Pijao Fault)
AC

Fault_2 58SE 1.35 x 10-4 1.38 x 10-5 1.86 x 10-9

Fault_3
82SE 4.43 x 10-6 2.49 x 10-6 1.10 x 10-11
(Manizales-Aranzazu Fault)

Fault_4
85NW 1.07 x 10-5 3.87 x 10-6 4.14 x 10-11
(San Jeronimo Fault)

17
ACCEPTED MANUSCRIPT

Fault_5
80SE 7.13 x 10-1 1.00 x 10-3 7.13 x 10-4
(Samaná Sur Fault)*

Fault_6
80SE 7.13 x 10-1 1.00 x 10-3 7.13 x 10-4
(Pico Terrible Fault)*
330 * Dip of Samaná Sur and Pico Terrible Faults were taken from the geological map of the NDR (Martínez et al., 2014).

PT
331 5.3 Laboratory analysis

332 The lithologic units identified at the sampling locations were the Quebradagrande Sedimentary

RI
333 Complex (Ksc), the Quebradagrande Volcanic Complex (Kvc), the Cajamarca Complex (Pes),

334 and the Andesite unit (NgQa; Table 1). Thin section analysis was used to define the rock type,

SC
335 characteristic texture and composition of the rock samples (Table 3).

336 The Quebradagrande Volcanic Complex (Kvc) shows evidence of regional metamorphism and

337
U
presence of hydrothermal activity. It is mainly composed of a fine-grained matrix (50%) due to
AN
338 the dynamic metamorphism and hydrothermal alteration, plagioclase (10%), olivine and/or

339 pyroxene (25%) and hydrothermal minerals like epidote (2.7%) close to the faults, for example in
M

340 the sample P3M2.


D

341 The Quebradagrande Sedimentary Complex (Ksc) is affected by dynamic metamorphism and it is

342 mainly composed of minerals with laminar fabric represented by fine micas featuring muscovite
TE

343 (49.5%), crystalline quartz (8.3%) and iron oxide type hematite (41.5%) close to faults.

344 According to thin section analysis, the Cajamarca Complex (Pes) has been affected by regional
EP

345 stress and temperature, and it contains evidence of dynamic metamorphism and hydrothermal

346 alteration like chalcedony mineral close to fault zones. The Cajamarca Complex (Pes) is mainly
C

347 constituted by quartz (48%) and muscovite (24.4%); to a lesser extent graphite (12.6%),
AC

348 plagioclase (1.4%), orthoclase (2.8%), chlorite (3.1%), and hematite (7.7%) were also identified.

349 Presence of chalcedony and calcite were observed in filled fractures.

350

351

352

18
ACCEPTED MANUSCRIPT

353

354

355

356 Table 3: Rock type characterization

PT
Code Lithologic unit Rock Type

Quebradagrande Sedimentary
P1M1 Sandstone

RI
Complex (Ksc)
Quebradagrande Volcanic Basalt or diabase with dynamic
P1M2
Complex (Kvc) metamorphism

SC
Quebradagrande Volcanic Basalt or diabase with dynamic
P2M1
Complex (Kvc) metamorphism

U
Quebradagrande Sedimentary
P3M1 Sandstone
Complex (Ksc)
AN
Quebradagrande Volcanic Basalt or diabase with cataclasis
P3M2
Complex (Kvc) and hydrothermal activity

Quebradagrande Volcanic Basalt or diabase with dynamic


M

P5M1
Complex (Kvc) metamorphism

Quebradagrande Volcanic Basalt or diabase with dynamic


P6M1
D

Complex (Kvc) metamorphism


Schist with dynamic and
P7M1 Cajamarca Complex (Pes)
TE

regional metamorphism
Schist with dynamic and
P8M1 Cajamarca Complex (Pes)
regional metamorphism
EP

Schist with dynamic and


P9M1 Cajamarca Complex (Pes)
regional metamorphism
P11M1 Andesite unit (NgQa) Andesite
357
C

358 Grain density, permeability, and hydraulic conductivity associated with each lithology were
AC

359 calculated as the mean value obtained from all the samples belonging to the same unit (Table 4).

360 The hydraulic conductivity of sample P7M1 (Pes) and P6M1 (Kvc) were higher than the other

361 samples of the same unit, which is probably due to the proximity to the San Jeronimo fault that is

362 the contact between these two complexes. These extreme values for Pes and Kvc would

19
ACCEPTED MANUSCRIPT

363 significantly affect the calculation of the mean. Hence, the median was used because it was

364 considered to better represent the hydraulic conductivity of these lithological units.

365 Samples 2014MI were collected during fieldwork conducted in 2014 in the same area. Some of

366 them were selected because of their good quality.

PT
367 Table 4: Physical and hydraulic properties measured in the laboratory parallel (p) and
368 perpendicular (v) to the foliation
369

RI
370 *Hydraulic conductivity of the Pes and Kvc units was calculated as the median of the sample values
Mean Mean
Grain
Lithological Sample Permeability grain hydraulic
density
unit code (m2) density conductivity

SC
(kg m-3)
(kg m-3) (m s-1)*
P1M1(p) 2480 -
P1M1(v) 2650 -
Ksc 2585 9.81 x 10-11

U
P3M1(p) 2600 -
P3M1(v) 2610 <9.87 x 10-18
AN
P1M2(v) 2610
P2M1(v) 2600 <9.87 x 10-18
Kvc P3M2(p) 2810 2724 9.81 x10-11
P3M2(v) 2930
M

P6M1(v) 2670 <9.87 x 10-16


NgQa 2014MI18 2660 <9.87 x 10-18 2660 2.02 x10-10
P7M1(p) 2550 -
D

P7M1(v) 2660 4.93 x 10-15


P8M1(p) 2660 -
P8M1(v) 2640 <9.87 x 10-18
TE

P9M1(p) 2660 -
Pes 2697 8.44 x 10-11
P9M1(v) 2670
2014MI17(p) 2790
EP

2014MI17(v) 2790 <9.87 x 10-18


2014MI21(p) 2780
2014MI21(v) 2770
371
C

372
373
AC

374 Several permeability data have the same value because most of our samples had permeability

375 lower than minimum range of detection (9.87 x 10-18 m2) of the gas permeameter used for the

376 measurement. In summary, the analysis reveals a rock matrix with a low permeability and

377 highlights the importance of fractures on fluid flow.

378

379 5.4 Geological cross-section

20
ACCEPTED MANUSCRIPT

380 The thickness of the lithologic units in the geological cross-section AA’ was based on a previous

381 cross-section of the NDR volcano published by Central Hidroelectrica de Caldas (CHEC et al.,

382 1983). The Andesite unit is approximately 500 m thick (CHEC et al., 1983) and lies above the

383 Cajamarca Complex. It is possible to see the different superficial deposits (Qto, Qar, Qfl)

PT
384 produced by the volcanic activity during the Quaternary period. The two members of the

385 Quebradagrande Complex (Ksc and Kvc) are located in the West of the cross-section and are

RI
386 dipping to the East. Fault dips are listed in Table 2.

U SC
AN
M
D
TE
C EP
AC

21
ACCEPTED MANUSCRIPT

387

388

PT
RI
SC
60° 60°
60°
60°

U
AN
M
389

D
390 Figure 6. Inferred geological cross-section AA’

TE
C EP
AC

22
ACCEPTED MANUSCRIPT

391 5.5 Numerical modeling

392 The simulated domain is based on the inferred geological cross-section AA’ (Figure 6), where

393 superficial geological deposits (Qar, Qfl, and Qto) are neglected, since their thickness is small

394 enough to have a negligible influence on the underground temperature due to their limited spatial

PT
395 extent and thickness. The zone of interest is situated around the Samaná Sur fault crossing the

396 Cajamarca Complex, which is the formation of most interest for future geothermal exploration

RI
397 (Almaguer, 2013). The triangular mesh built with GMSH has 72663 elements, with an average

SC
398 mesh element size of 100 m in most of the domain, and 50 m close to the top and to the faults to

399 properly capture the topographic variations and the influence of faults. Groundwater flow

U
400 boundary conditions are hydraulic heads equal to the topographic elevation at the lateral

401 extremities and at the top of the model, while the bottom boundary is considered impermeable.
AN
402 The top heat transfer boundary condition is a constant temperature determined from an elevation-

403 dependent temperature profile, varying from 18 to 4 °C (CORPOCALDAS, 2005). Adiabatic


M

404 conditions are set to the left and right of the domain. The bottom heat transfer boundary is a
D

405 linearly variable heat flux defined considering the maximum, the minimum and the mean heat
TE

406 flow values of the study area. Hamza et al. (2005) through polynomial methods examined heat

407 flow large-scale variations overt the South America continent and reported a minimum value of
EP

408 heat flux in the area of NDR of 64 mW m-2. However, for the modeling scale considered in this

409 work, it is more appropriate to use the data available from geothermal gradient wells drilled in
C

410 2011 (Rojas, 2012) and from the Colombian heat flow map. The minimum value, 120 mW m-2, is
AC

411 equal to the smallest heat flow in the study area, according to the Colombian heat flow map

412 (INGEOMINAS, 2000). The mean value, 178 mW m-2, is evaluated from the geothermal

413 gradient measured in the 300 m deep wells presented by Rojas (2012); the maximum value, 270

414 mW m-2, is estimated considering the distance from the NDR peak, where the maximum heat

415 flow of 366 mW m-2 was evaluated from the temperature gradient observed at the Nereidas well

416 (Bernal et al., 2000).

23
ACCEPTED MANUSCRIPT

417 The total simulated time was 10 million years: 10 time steps of 1 million years were considered

418 for transient heat transfer, while groundwater flow was simulated in steady-state condition. The

419 heat source given by the volcanic activity in the area affects the heat transfer process. Time steps

420 of the order of magnitude of geological periods are needed to reproduce the current temperature

PT
421 distribution at depth. However, there are no changes in the simulated temperature after the sixth

422 time step, since near steady-state conditions are already reached at that time.

RI
U SC
AN
M

423
D

424 Figure 7. Model extension and boundary conditions for the numerical model.

425 The thermal properties measured by Vélez et al. (2018), grain density, and hydraulic conductivity
TE

426 calculated from permeability were used as inputs for the numerical simulations (Tables 2 and 5).

427 It is known that rock properties vary with temperature (Whittington et al., 2009; Nabelek et al.
EP

428 2012), as shown in previous modeling studies on the NDR area (Vélez et al., 2018). Nevertheless,
C

429 the focus of this work is to evaluate the influence of faults and faults dip on the groundwater flow

430 and heat transfer processes, such that constant thermal properties are considered to simplify the
AC

431 model and to reduce the computational effort. The hydraulic conductivity of Cajamarca Complex

432 (Pes) was 8.44 x10-11 m s-1. However, the field value can be higher than laboratory measurements

433 considering the intensity of the regional metamorphism (Monsalve et al., 1998; González, 2001;

434 Almaguer, 2013) that cannot be represented at the scale of the rock sample having an average

24
ACCEPTED MANUSCRIPT

435 area of 250 cm2. Thus, a value of 8.44 x 10-10 was selected for the numerical simulations. Thermal

436 properties of Kvc were not available and values equal to those of Ksc were assumed.

437 Table 5. Properties of porous medium used for numerical simulations


Properties Pes NgQa Kvc Ksc

PT
Hydraulic conductivity K (m s-1) 8.44 x10-10 2.02 x10-10 9.81 x10-11 9.81 x10-11

Thermal conductivity (W m-1 K-1) 2.98 1.22 1.92 2

RI
Specific heat capacity (J kg-1 K-1) 910 815 830 815

Density (kg m-3) 2697 2660 2724 2585

SC
438

439 Cherubini et al. (2013) described that pressure and temperature variations (up to 5°C) can occur

U
440 between a model with a dipping fault of 60° and a model with a vertical fault. They point out the
AN
441 importance of the implementation of inclined faults, which is a closer approximation to the

442 natural fault geometry in the model. Taking into account their results, three modeling scenarios
M

443 were defined (Table 6) to investigate the influence of fault dip with respect to the groundwater

444 flow and heat transfer processes. The Samaná Sur fault was selected for simulations because it
D

445 crosses the geothermal reservoir (Pes) and it is located outside of the Los Nevados NNP, where
TE

446 future geothermal exploitation can take place. The first case is characterized by an 80° SE dip

447 (CASE A) reported in the geologic map of the NDR (Martínez et al., 2014), the second case by a
EP

448 dip of 45° SE (CASE B), and the third one by a dip of 80° SW (CASE C). Although there is a

449 geological map available for the study area, these modeling scenarios were defined because there
C

450 is still a lot of uncertainty about fault dip and down-dip continuity of faults, which justifies the
AC

451 use of alternative scenarios.

452

453

454

455

25
ACCEPTED MANUSCRIPT

456 Table 6. Simulated temperature and hydraulic gradient at the Samaná Sur fault for the three
457 modeling scenarios considered (CASE A, B, and C)
Simulated
Dip of Hydraulic Velocity along the
temperature
Scenario Samaná Sur Gradient along the Samaná Sur fault
at surface
Fault (°) Samaná Sur fault (m s-1)
(°C)
CASE A 80SE 60.5 0.00108 7.70 x10-4

PT
CASE B 45SE 68.1 0.00220 1.57 x10-3

CASE C 80SW 54 0.00093 6.63 x10-4

RI
458

U SC
AN
M
D
TE
C EP
AC

459
460 Figure 8. Velocity vectors considering dip 80° SE for the Samaná Sur fault (CASE A)

26
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

461
462 Figure 9. Simulated temperature distribution considering dip 80° SE (CASE A), 45° SE (CASE

463 B) and 80° SW (CASE C) of Samaná Sur fault

27
ACCEPTED MANUSCRIPT

464 6 Discussion

465 6.1 Groundwater flow

466 The hydraulic head applied to the lateral and top boundaries of the model generates a regional

467 flow from East to West. During the simulations, the hydraulic conductivity of the porous

PT
468 medium, aperture, transmissivity, hydraulic conductivity, and hydraulic head condition at the top

469 of the fault are the parameters that cause the fault to behave as a pathway or barrier to the

RI
470 transport of fluid and heat. Varying any of these parameters results in relevant differences in the

SC
471 pressure, velocity, and temperature field.

472 According to Figure 8 and Figure 9, the Samaná Sur fault seems to be the fault that has the

highest impact on the hydrothermal process. The high transmissivity value (7.13 x 10-4 m2 s-1) and

U
473

474 the hydraulic conductivity contrast between the fault (7.13 x 10-1 m s-1) and the surrounding
AN
475 porous medium (8.44 x 10-10 m s-1) allows this fault to behave as a preferential pathway. The

476 outflow observed at this fault depends on the regional East-West groundwater circulation and on
M

477 the low hydraulic head at its top, since the fault´s up-dip termination is located in a valley.
D

478 Although the other faults represent an outflow of fluid, they are not permeable enough or do not
TE

479 have enough hydraulic head contrast to induce a remarkable influence on the fluid circulation and

480 heat transfer. The Samaná Sur and Pico Terrible faults have the same hydraulic conductivity,
EP

481 aperture and transmissive conditions, representing highly permeable structures. However, since

482 the Samaná Sur fault crosses the topography in a valley at lower elevation than the surroundings
C

483 areas, it creates a significant hydraulic head difference that induces fluid flow and heat transfer
AC

484 along the fault. On the contrary, the up-dip termination of the Pico Terrible fault is located at high

485 elevations that do not generate a significant hydraulic gradient to influence the flow circulation.

486 Similar results were found by Cherubini et al. (2013) and Cherubini et al. (2014), who mentioned

487 that the variation of the fault permeability or the width of the fault zone affects the entire

488 pressure, velocity, and temperature solution of the fractured numerical model. They additionally

28
ACCEPTED MANUSCRIPT

489 found a minor effect on the heat transport along the fault with the lowest fault zone width due to

490 the smaller volume of water being transported.

491

492 6.2 Heat Transport

PT
493 The maximum temperature simulated for CASE A with coupled groundwater flow and advective

494 heat transfer (Figure 9) was 520 °C at the bottom right-hand side of the model, where the largest

RI
495 heat flux (270 mW m-2) was applied as a bottom boundary condition. High influence of Samaná

SC
496 Sur fault was observed in all cases.

497 The hydraulic gradient for CASES A and C is similar because the length of the fault is the same

U
498 and the hydraulic head at its bottom in CASE A is only slightly higher than in CASE C. The

499 CASE B is characterized by the highest hydraulic gradient because, even if the length of the fault
AN
500 is larger than in the other two cases, the hydraulic head at the bottom of the fault is considerably

501 higher. The simulated temperature at the surface, where the Samaná Sur fault crosses the
M

502 topography, is 60.5 °C for CASE A, which is similar to the measured temperature at the Hotel
D

503 Termales hot spring (63°C), closely located to this fault. The simulated temperature at the same
TE

504 location for CASE B and CASE C is 68.1 °C and 55°C, respectively, showing the influence of

505 fault dip and trend in the hydraulic gradient and, as a consequence, in the transport of hot fluids.
EP

506 When the trend is perpendicular and dip is in favor of the fluid flow (< 90° facing fluid flow) as

507 in CASE A and CASE B, the hydraulic gradient is higher and the simulated surface temperature
C

508 increases. On the other hand, if the trend is perpendicular, but the dip of fault is opposite to the
AC

509 fluid flow (> 90° facing fluid flow) as in CASE C, the hydraulic gradient is lower and the

510 simulated surface temperature decreases. The reproduction of the temperature measured at the

511 Hotel Termales hot spring supports the suitability of the methodology applied for this work. The

512 bottom of the fault with the lowest dip (45° SE; CASE B) is closer to the heat source and it has a

513 greater hydraulic gradient than CASE A and CASE C, resulting, as expected, in the highest

514 simulated temperature at the surface. Additionally, according to Kalinina et al. (2013) and Riahi

29
ACCEPTED MANUSCRIPT

515 et al. (2014), heat transport through faults is greatly controlled by the distance traveled by the

516 fluid. A longer path result in slower temperature decay and horizontal fractures are the best cases

517 for the heat transport and heat extraction, since a larger reservoir volume is involved in the heat

518 exchange. This behavior is similar to what happens in CASE B, where the longest path is

PT
519 observed, resulting in a larger reservoir interaction and the maximum simulated temperature at

520 the surface. Although the variation of surface temperature for cases A, B and C seems small, this

RI
521 result is important for geothermal exploration since it demonstrates that the dip of faults affects

SC
522 the temperature of the fluid at surface. The small variation may be related tothe discrete fracture

523 conceptual model used to represent the faults since real faults systems are expected broader.

U
524

525 6.3 Geothermal exploration


AN
526 The numerical results obtained here provide important information for further geothermal

527 exploration in the NDR area and in similar geothermal systems with hydraulic gradient driven by
M

528 topography and faults zones affecting groundwater circulation and heat transfer.
D

529 Low-permeability fault zones seem to induce no remarkable effects on the temperature
TE

530 distribution. Therefore, those zones are not recommended as a target for geothermal exploration

531 well drilling. Additionally, faults located beyond San Jeronimo Fault are too far from the
EP

532 expected geothermal reservoir and the heat source, hence it would be difficult to find a good

533 geothermal resource at an economically favorable drilling depth in that area.


C

534 Drilling a geothermal well close to a permeable fault zone would be recommended because these
AC

535 zones enhance flow of hot and deep water through the fault. In this work, the fault that constitutes

536 the best target is the Samaná Sur Fault, which is close to the heat source (Nevado del Ruiz peak)

537 outside of the Los Nevados NNP and crossing the Cajamarca Complex, the geological formation

538 of most interest. Meeting these conditions, the development of a high-enthalpy geothermal project

539 would be feasible, although numerical modeling can be improved by additional field based

30
ACCEPTED MANUSCRIPT

540 studies (geophysical, structural, and geochemical), to better delineate the zone of interest before

541 starting any exploration drilling activity.

542

543 6.4 Model limitation

PT
544 The main advantage of the DFM applied here is to represent the faults as discrete features with

545 their own geometry and elements embedded in the porous matrix. These elements, represented as

RI
546 1D-lines, are assumed to represent open and saturated fractures. A refined mesh is usually

SC
547 required to represent these discrete features with physical, hydraulic and thermal properties

548 different from those of the surrounding porous medium.

U
549 Vélez et al. (2018) conducted sensitivity tests and suggested hydraulic conductivity was one of

550 the most important parameters affecting simulated temperature. For faults crossing the
AN
551 topography in a depression, a hydraulic contrasts between the fault itself and the surrounding area

552 is created, because of the flow boundary condition specified to the top boundary of the model
M

553 (head = topographic elevation, as shown in Figure 7). This hydraulic contrast and the hydraulic
D

554 conductivity of the porous media and of the faults are the parameters that have most impact on
TE

555 the simulated temperature.

556 The few structural and hydraulic data available in the study area are the main limitation of the
EP

557 model. It is therefore important to account for more structural and geological data, such as the

558 aperture, spacing, dip angle, and fracture filling material, to build more accurate 2D models and
C

559 perform new 3D models around faults.


AC

560 This study is the first attempt to investigate the impact of fault zones on groundwater flow and

561 heat transfer in the area located Northwest of the Nevado del Ruiz volcano. It represents an

562 important tool that can be updated as soon as new data are available to support geothermal

563 exploration in the area.

564

565

31
ACCEPTED MANUSCRIPT

566 7 Conclusions

567 This paper presented a characterization of faults to the Northwest of the Nevado Del Ruiz (NDR)

568 volcano (Colombia). The work consisted of collecting rock samples at surface outcrops, defining

569 their rock type and composition, measuring hydraulic properties of porous and fractured medium

PT
570 using laboratory and field methods, inferring a geological cross-section and simulating different

571 heat transfer modeling scenarios. The main goal of this modeling exercise was to understand the

RI
572 influence of faults in the transport of hot fluids in the geothermal reservoir, which is a critical

SC
573 aspect for the development of geothermal projects in a site like the NDR volcano.

574 The accurate match between simulated (CASE A, with the real fault dip) and observed

U
575 temperature (measured at the Hotel Termales hot spring) demonstrates the validity of the

576 numerical simulation approach considered here.


AN
577 According to the analysis of shallow seismic data, most of the earthquake alignments correspond

578 to the fault locations, which is the evidence of the fault activity. The intersection of faults may
M

579 indicate the location of recent and high permeability zones that may increase hot fluid circulation.
D

580 Taking into account the simulation results, it is possible to conclude that fault dip and trend have
TE

581 an influence on the hydraulic gradient affecting fluid flow and heat transport toward the surface

582 of the system. When the dip is in favor of the fluid flow (< 90° facing fluid flow), like in CASE A
EP

583 and CASE B, the simulated surface temperature increases. On the other hand, if the dip of fault is

584 opposite to the fluid flow (> 90° facing fluid flow), like in CASE C, the simulated surface
C

585 temperature decreases. Additionally, it is possible to conclude that CASE B has the highest
AC

586 simulated surface temperature since the fault is closer to the heat source, it has a greater hydraulic

587 gradient and a longer flow path, which means a larger water-rock interaction.

588 Findings reported in this paper support the need for more structural geology studies where the

589 presence of tectonic syntax and earthquake alignment intersections exists, in order to better

590 characterize and understand fault behavior and to improve future numerical models. 3D

591 geomodeling coupled to numerical simulations considering two-phase (liquid and vapor) fluids,

32
ACCEPTED MANUSCRIPT

592 with temperature-dependent water properties, and combined with new field data will further

593 improve the current understanding of the NDR hydrothermal system to the benefit of geothermal

594 exploration. Geomodeling tools can be particularly useful to better characterize the faults and

595 geological contacts in 3D space. The Cajamarca Complex appears highly heterogeneous and

PT
596 detailed field investigations are required to improve its description.

597 The first type boundary condition, which creates a contrast between heads at the fault nodes and

RI
598 in the surrounding porous medium, as well as the hydraulic conductivity, were the most important

SC
599 elements influencing the groundwater flow and heat transfer mechanism, as shown by the high

600 sensitivity of the modeling results to their variation. Water kinematic viscosity depends on the

U
601 temperature and was required to calculate the hydraulic conductivity from permeability.

602 Therefore the temperature considered for this calculation had an impact on the simulated
AN
603 temperature. Additional accurate hydraulic conductivity measurements on core samples can help

604 to get better estimate of heat transport and temperature at depth. Such logical steps are needed to
M

605 move forward with the development of geothermal energy in Colombia, providing critical
D

606 knowledge to support geothermal exploration and energy decisions. These recommendations also
TE

607 apply for other countries of South America, where there are ongoing characterization and

608 evaluation of geothermal resources (Bertani, 2016).


EP

609

610 8 Acknowledgments
C

611 This work has been conducted in the context of a research project funded by the Universidad de
AC

612 Medellín (Medellín, Colombia) in collaboration with the Institut National de la Recherche

613 Scientifique (Québec, Canada). Thanks to the Canadian Department of Foreign Affairs, Trade

614 and Development who assigned the Emerging Leaders in the Americas Program scholarships to

615 David Alejandro Moreno Rendón in 2016.

616 The IGCP, UNESCO, and IUGS are finally acknowledged since this work is part of the project

617 “IGCP 636-Unifying international research forces to unlock and strengthen geothermal

33
ACCEPTED MANUSCRIPT

618 exploitation of the Americas and Europe”, currently funded by UNESCO (United Nations

619 Educational, Scientific and Cultural Organization) and IUGS (International Union of Geological

620 Sciences) within the International Geoscience Programme (IGCP).

621 References

PT
622 Alfaro, C. 2015. Improvement of perception of the geothermal energy as a potential source of
623 electrical energy in Colombia, country update. Paper presented at the World Geothermal
624 Congress, Melbourne, Australia, p. 15.

RI
625 Almaguer, J.L. 2013. Estudios magnetotelúrico con fines de interés geotérmico en sector Norte
626 del Nevado del Ruíz, Colombia. MSc thesis, Universidad Nacional Autónoma de México, p. 139.

SC
627 American Petroleum Institute. 1998. Recommended Practices for Core Analysis (Vol. 2), p. 236.
628 Arango, E.E., Buitrago, A.J., Cataldi, R., Ferrara, G.C., Panichi, C., Villegas, V.J. 1970.
629 Preliminary study on the Ruiz geothermal project (Colombia). Geothermics, 2, Part 1: 43-56. doi:

U
630 10.1016/0375-6505 (70)90005-2.
AN
631 Aravena, D., Muñoz, M., Morata, D., Lahsen, A., Parada, M. Á., Dobson, P. 2016. Assessment of
632 high enthalpy geothermal resources and promising areas of Chile. Geothermics, 59, Part A, 1–13.
633 doi: 10.1016/j.geothermics.2015.09.001.
M

634 Bakhsh K., Nakagawa M., Arshad M., Dunnington L. 2016. Modeling Thermal Breakthrough in
635 Sedimentary Geothermal System, Using COMSOL Multiphysics. Proceedings 41st Workshop on
636 Geothermal Reservoir Engineering Stanford University, Stanford, California, February 22-24.
637 SGP-TR-209, p. 11.
D

638 Benavente O., F. Tassi, M. Reicha, F. Aguilera , F. Capecchiacci, F. Gutiérrez , O. Vaselli, A.


TE

639 Rizzo. 2016. Chemical and isotopic features of cold and thermal fluids discharged in the Southern
640 Volcanic Zone between 32.5°S and 36°S: Insights into the physical and chemical processes
641 controlling fluid geochemistry in geothermal systems of Central Chile. Chemical Geology 420:
642 97-113.
EP

643 Bernal N.F., Ramirez G., Alfaro C.V. 2000. Mapa geotérmico de Colombia. Versión 1.0. Escala
644 1:1’500.000. Memoria explicativa. Exploración y Evaluación de Recursos Geotérmicos. Instituto
645 de investigación e información geocientífica, minero-ambiental y nuclea INGEOMINAS, p. 51.
C

646 Bertani, R. 2016. Geothermal power generation in the world 2010-2014 update report.
AC

647 Geothermics, 60: 31-43. doi:http://dx.doi.org/10.1016/j.geothermics.2015.11.003.


648 Böttcher, N., Watanabe, N., Görke, U.-J., Kolditz, O. 2016. Geoenergy Modeling I. Geothermal
649 Processes in Fractured Porous Media. SpringerBriefs in Energy. Computational Modeling of
650 Energy Systems, p. 117.
651 Caine, J. S., Evans, J. P., Forster, C. B. 1996. Fault zone architecture and permeability structure.
652 Geology, 24(11), 1025-1028.

34
ACCEPTED MANUSCRIPT

653 Camargo, G., Mojica, J. 2004. Reactivation (long-term evolution) of Silvia-Pijao Fault along the
654 "Quebrada La Maizena" western flank of Central Cordillera, Quindio-Colombia, Geologia
655 Colombiana, 29, 11-22.
656 Ceballos, D. 2017. Análisis geológico y estructural detallado de una zona del proyecto
657 geotérmico en el valle de Las Nereidas, macizo volcánico Nevado del Ruiz, para contribuir en el
658 proceso de exploración geotérmica, CHEC, Manizales: Universidad de Caldas. p. 70.

PT
659 CHEC - Central Hidroeléctrica de Caldas, Instituto Colombiano de Energía Eléctrica, Consultoría
660 Técnica Colombiana Ltda, Geotérmica Italiana. 1983. Investigación Geotérmica Macizo
661 Volcánico del Ruíz. Fase II, Etapa A (Vol. II, III). Bogotá, p. 194.

RI
662 Cherubini, Y., Cacace, M., Blocher, G., Scheck-Wenderoth, M. 2013. Impact of single inclined
663 faults on the fluid flow and heat transport: results from 3-D finite element simulations.
664 Environmental Earth Sciences.70: 3603–3618.

SC
665 Cherubini, Y., Cacace, M., Scheck-Wenderoth, M., Noack, V. 2014. Influence of major fault
666 zones on 3-D coupled fluid and heat transport for the Brandenburg region (NE German Basin).
667 Geothermal Energy Science.2, 1-20.
668
U
Colombani J., Lamagat J.-P., Thiebaux, J. 1972. Mesure de la perméabilité des sols en place: un
AN
669 nouvel appareil pour la méthode Muntz une extension de la méthode Porchet aux sols
670 hétérogènes. Cah. O.R.S.T.O.M., sér. Hydrol., IX(3), p. 31.
671 Corbel, S., Schilling, O., Horowitz, F., Reid, L., Sheldon, H., Timms, N., Wilkes, P. 2012.
M

672 Identification and Geothermal Influence of Faults in the Perth Metropolitan Area, Australia.
673 Workshop on Geothermal Reservoir Engineering. Stanford: Stanford University, Stanford
674 Geothermal Program, p. 8.
D

675 CORPOCALDAS 2005. Estudio sobre el estado actual de los páramos del Departamento de
676 Caldas: Línea Base. Conservación Internacional Colombia. p. 299.
TE

677 El Espectador. 2014. Volcán Nevado del Ruíz es el más vigilado del mundo.
678 https://www.elespectador.com/noticias/nacional/volcan-nevado-del-ruiz-el-mas-vigilado-
679 delmundo-articulo-527469.
EP

680 Ferrill, D., Sims, D., Waiting, D., Morris, A., Franklin, N., Schultz, A. 2004. Structural
681 framework of the Edwards Aquifer recharge zone in south-central Texas. GSA Bulletin 116 (3/4):
682 406-418.
C

683 Forero, J.A. 2012. Caracterización de las alteraciones hidrotermales en el flanco Noroccidental
AC

684 del Volcán Nevado del Ruiz, Colombia. MSc thesis, Universidad Nacional de Colombia, Bogotá,
685 p. 121.
686 Geuzaine, C., Remacle, J-F. 2009. A three-dimensional finite element mesh generator with built-
687 in pre- and post-processing facilities. International Journal for Numerical Methods in Engineering
688 79(11): 1309-1331.
689 Gibson, R. 1998. Physical character and fluid-flow properties of sandstone derived fault zones.
690 Geological Society, London, Special Publications, 127, 83-97.
691 https://doi.org/10.1144/GSL.SP.1998.127.01.07.

35
ACCEPTED MANUSCRIPT

692 González, H. 2001. Geología de las planchas 206 Manizales y 225 Nevado del Ruíz. Memoria
693 explicativa. Instituto de investigación e información geocientífica, minero-ambiental y nuclear,
694 INGEOMINAS, Bogotá, p. 93.
695 González, H. 1980. Geología de las planchas 167 (Sonson) y 187 (Salamina). INGEOMINAS,
696 23(1), p. 174.
697 González-García, J., Hauser, J., Annetts, D., Franco, J., Vallejo, E., Regenauer-Lieb, K. 2015.

PT
698 Nevado Del Ruiz volcano (Colombia): a 3D model combining geological and geophysical
699 information. World Geothermal Congress, Melbourne, Australia, p. 11.
700 González-García J., Jessell M. 2016. A 3D geological model for the Ruiz-Tolima Volcanic

RI
701 Massif (Colombia): Assessment of geological uncertainty using a stochastic approach based on
702 Bézier curve design. Tectonophysics 687, 139–157.

SC
703 Gómez, J., Montes, N.E., Nivia, Á., Diederix, H., Compiladores. 2015. Plancha 5-09 del Atlas
704 Geológico de Colombia 2015. Escala 1:500 000. Servicio Geológico Colombiano. Bogotá.
705 Hamza V.M., Silva Dias F.J.S., Gomes, A.J.L., Delgadilho Terceros Z.G. 2005. Numerical and

U
706 functional representations of regional heat flow in South America. Physics of the Earth and
707 Planetary Interiors. 152, 223–256.
AN
708 Hao, Y., Fu, P., Johnson, S.M., Carrigan, C.R. 2012. Numerical Studies of Coupled Flow and
709 Heat Transfer Processes in Hydraulically Fractured Geothermal Reservoirs. GRC Transactions
710 36, 453-458.
M

711 Holzbecher E., Wong L.W., Litz M.-S. 2010. Modelling Flow through Fractures in Porous Media.
712 Proceedings of the COMSOL Conference, Paris, p. 7.
D

713 INGEOMINAS 2000. Mapa de flujos de calor.


714 Jaffre, J., Mnejjab, M., Roberts, J. 2011. A discrete fracture model for two-phase flow with
TE

715 matrix-fracture interaction. Procedia Computer Science. 4, 967-973.


716 Kalinina, E., Klise, K., McKenna, S., Hadgu, T., Lowry, T. 2013. Applications of the fractured
717 continuum model (FCM) to EGS heat extraction problems. Thirty-Eighth Workshop on
EP

718 Geothermal Reservoir Engineering. Stanford, California: Stanford University, p. 16.


719 Li, W., Yost, K., Sousa, R. 2013. Heat Transfer Between Fluid Flow and Fractured Rocks.
720 Massachusetts Institute of Technology . GRC Transaction 37, 165-171.
C

721 Londoño, J., Sudo, Y. 2002. Velocity structure and a seismic model for Nevado del Ruiz Volcano
AC

722 (Colombia). Journal volcanology and geothermal research, 119, 61-87.


723 Lovelessa, S., Pluymaekers, M., Lagroua, D., De Boeverc, E., Doornenbalb, H., Laenena, B.
724 2014. Mapping the geothermal potential of fault zones in the Belgium-Netherlands border region.
725 European Geosciences Union General Assembly. Energy procedia, 351-358.
726 Martínez, L., Valencia, L., Ceballos, J., Narváez, B., Pulgarín, B., Correa, A., Pardo, N. 2014.
727 Geología y Estratigrafía del Complejo Volcánico Nevado del Ruiz. Bogotá-Popayán-Manizales:
728 Servicio Geológico Colombiano, p. 853.

36
ACCEPTED MANUSCRIPT

729 Mejía, E., Velandia, F., Zuluaga, C., López, J., Cramer, T. 2012. Análisis estructural al noreste
730 del Volcán Nevado del Ruíz, Colombia – Aporte a la exploración geotérmica. Boletín de
731 Geología, 34(1), 1-15.
732 Monsalve, M.L., Rodriguez, G.I., Mendez, R.A., Bernal, N.F. 1998. Geology of the Well
733 Nereidas 1, Nevado Del Ruiz Volcano, Colombia Geothermal Resources Council 22, 1-6.
734 Mosquera, D., Marín, P., Vesga, C., González, H., 1998a. Geología de la Plancha 225. Escala

PT
735 1:100 000. Nevado del Ruíz.
736 Mosquera, D., Marín, P., Vesga, C., González, H., Maya, M., 1998b. Geología de la Plancha 206.
737 Escala 1:100 000. Manizales.

RI
738 Nabelek, P. I., Hofmeister, A. M., Whittington, A. G. 2012. The influence of temperature-
739 dependent thermal diffusivity on the conductive cooling rates of plutons and temperature-time

SC
740 paths in contact aureoles. Earth and Planetary Science Letters, 317, 157-164.
741 Oosterbaan, R.J. Nijland, H.J. 1994. Determining the saturated hydraulic conductivity. In H.P.
742 Ritzema, ed. Drainage principles and applications. 2nd edition. ILRI Publication 16. Wageningen,

U
743 The Netherlands, ILRI, 435-475.
744 Pérez, P., Sánchez, P., Arancibia, G., Cembrano, J., Veloso, E., Lohmar, S., Stimac, J., Reich, M.,
AN
745 Rubilar, J., 2012. Sampling and detailed structural mapping of veins, fault-veins and faults from
746 Tolhuaca Geothermal System, southern Chile. XIII Congreso Geológico Chileno, Universidad
747 Catolica del Norte, Antofagasta, Chile. 495-497.
M

748 Puga-Lagos, P. 2012. Estudio experimental del Coeficiente de Permeabilidad en Arenas. Trabajo
749 de Grado, Universidad Católica de la Santísima Concepción, Facultad de Ingeniería,
750 Departamento de Ingeniería Civil, p. 189.
D

751 Riahi, A., Furtney, J., Damjanac, B. 2014. Evaluation of Optimum Well Positioning in Enhanced
TE

752 Geothermal Reservoirs Using Numerical Modeling. Itasca Consulting Group, Inc. Minneapolis,
753 Minnesota, U.S.A. Geothermal Resources Council, 38, 325-330.
754 Rojas, O.E. 2012. Contribución al modelo geotérmico asociado al sistema volcánico Nevado del
755 Ruiz-Colombia, por medio del análisis de la relación entre la susceptibilidad magnética,
EP

756 conductividad eléctrica y térmica del sistema. MSc thesis, Universidad Nacional de Colombia,
757 Bogotá, p. 183.
C

758 Sanchez-Alfaro P., Sielfeld G., VanCampen B., Dobson P., Fuentes V., Reed A., Palma-Behnke
759 R., Morata D. 2015. Geothermal barriers, policies and economics in Chile – Lessons for the
AC

760 Andes. Renewable and Sustainable Energy Reviews 51, 1390–1401.


761 Spence, W., Sipkin, S., Choy, G. 1989. Colorado, USA: U.S. Geological Survey. Earthquakes
762 and Volcanoes.Vol. 21, p. 64.
763 Stix, J., Layne, G., Williams, S. 2003. Mechanisms of degassing at Nevado del Ruiz volcano,
764 Colombia. Journal of the Geological Society, 160, 507-521.
765 Thouret, J.C. 1988. Les Andes Centrales de Colombie et leurs bordures; morphogénèse plio-
766 quaternaire et dynamique actuelle et récente d’une cordillère volcanique englacée. Univ. J.
767 Fourier, Grenoble, Thèse d’état, p. 628.

37
ACCEPTED MANUSCRIPT

768 Thouret, J.C. 1990. Effects of the November 13, 1985 eruption on the snow pack and ice cap of
769 Nevado del Ruiz volcano, Colombia. Journal of Volcanology and Geothermal Research, 41 (1),
770 177-201. doi:http://dx.doi.org/10.1016/0377-0273 (90)90088-W.
771 Toro, R., Osorio, J. 2005. Determinación de los tensores de esfuerzos actuales para el segmento
772 norte de los andes calculados a partir de mecanismos focales de sismos mayores. 27(44), 1-12.
773 Trenkamp, R., Kellogg, J.N., Freymueller, J.T., Mora, H.P. 2002. Wide plate margin deformation,

PT
774 southern Central America and Northwestern South America, CASA GPS observations. Journal of
775 South American Earth Sciences, 15 (2), 157–171. doi:http://dx.doi.org/10.1016/S0895-9811
776 (02)00018-4.

RI
777 Van der Molen, W., Martínez, J., Ochs, W. 2007. Guidelines and computer programs for the
778 planning and design of land drainage systems. FAO, p. 233.

SC
779 Vélez M.I., Blessent D., López I.J., Raymond J., Parra E. 2018. Geothermal potential assessment
780 of the Nevado del Ruiz volcano based on rock thermal conductivity measurements and risk
781 analysis correction. Journal of South American Earth Sciences 81, 153-164.

U
782 Witherspoon PA, Wang JSY, Iwai K, Gale JE. 1980. Validity of cubic law for fluid flow in a
783 deformable rock fracture. Water Resources Research, 16, 1016–1024
AN
784 Whittington, A. G., Hofmeister, A. M., Nabelek, P. I. 2009. Temperature-dependent thermal
785 diffusivity of the Earth’s crust and implications for magmatism. Nature Journal, 458, 319-321.
M
D
TE
C EP
AC

38
ACCEPTED MANUSCRIPT

Highlights

• Characterization of an area of geothermal interest in Colombia


• Heat-transfer modeling along inclined fault zones
• 2D numerical modeling of coupled groundwater flow and heat transfer
• Rock sampling and laboratory characterization of physical and thermal properties

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like