You are on page 1of 12

Journal of Volcanology and Geothermal Research 345 (2017) 172–183

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

journal homepage: www.elsevier.com/locate/jvolgeores

Stable isotopes evidence of recycled subduction fluids in the


hydrothermal/volcanic activity across Nicaragua and Costa Rica
Ramírez-Leiva A. a, Sánchez-Murillo R. a,⁎, Martínez-Cruz M. b, Calderón H. c, Esquivel-Hernández G. a,
Delgado V. d, Birkel C. e,f, Gazel E. g, Alvarado G.E. h, Soulsby C. f
a
Stable Isotope Research Group, Chemistry School, Universidad Nacional, P.O. Box: 86-3000, Heredia, Costa Rica
b
Observatorio Sismológico y Vulcanológico de Costa Rica, Universidad Nacional, P.O. Box: 2346-3000, Heredia, Costa Rica
c
Institute of Geology and Geophysics, IGG-CIGEO, UNAN-Managua, P.O. Box: 4598, Nicaragua
d
Centro para la Investigación en Recursos Acuáticos de Nicaragua, CIRA, UNAN-Managua, P.O. Box: 4598, Nicaragua
e
Department of Geography, University of Costa Rica, P.O. Box 11501-2060, San José, Costa Rica
f
Northern Rivers Institute, University of Aberdeen, P.O. Box: AB24 3UF, Aberdeen, Scotland
g
Earth and Atmospheric Science Department, Cornell University, P.O. Box: 14853-1504, Ithaca, NY, USA
h
Instituto Costarricense de Electricidad, P.O. Box: 10032-1000, San José, Costa Rica

a r t i c l e i n f o a b s t r a c t

Article history: The Central America volcanic front provides a unique opportunity to study hydrothermal inputs and their interac-
Received 10 July 2017 tion and mixing with modern meteoric waters. The objectives of this study were to: a) characterize the isotopic
Received in revised form 27 August 2017 composition (δ18O, δ2H, d-excess, and lc-excess) of hydrothermal/volcanic systems, b) analyze the influence of ki-
Accepted 28 August 2017
netic fractionation and meteoric water inputs in the isotopic composition of hydrothermal waters, and c) estimate
Available online 1 September 2017
the ‘andesitic water’ contribution (recycled subduction fluids) within the volcanic front of Nicaragua and Costa
Keywords:
Rica. Hydrothermal evaporation lines are described as: δ2H = 4.7·δ18O − 13.0 (Costa Rica) and δ2H = 2.7·δ18O
Central America volcanic front − 31.6 (Nicaragua). These regressions are significantly (p b 0.001) deviated from their respective meteoric
Hydrothermal systems water lines: δ2H = 7.6·δ18O + 7.4 (Costa Rica) and δ2H = 7.4·δ18O + 5.2 (Nicaragua). The greater rainfall inputs
Isotopic fractionation in Costa Rica with respect to Nicaragua, resulted in the attenuation of the evaporative effect as observed in the
Recycled subduction fluids strong bimodal distribution of the hydrothermal waters, which can be divided in fluids: a) isotopically-close to me-
Groundwater mixing teoric conditions and b) isotopically-altered by the interaction with recycled subduction fluids and kinetic fraction-
ation. The latter is clearly depicted in the significantly (p b 0.001) low d-excess and lc-excess median values
between Costa Rica (+5.10‰, −5.25‰) and Nicaragua (−2.42‰, −10.65‰), respectively. Poor correlations be-
tween δ18O/δ2H and the elevation gradient emphasize that the contribution of recycled subduction fluids and sub-
sequent surface kinetic fractionation are the main drivers of the isotopic departure from the orographic distillation
trend captured in the rainfall isoscapes. End-member mixing calculations resulted in a significant difference (p b
0.001) between the mean ‘andesitic water’ contribution to the hydrothermal systems of 15.3 ± 10.8 (%, ±1σ)
(Nicaragua) and 19.7 ± 10.3 (%, ±1σ) (Costa Rica). The spectrum of ‘andesitic water’ contribution largely reflects
the degree of mixing with isotopically ‘pre-shifted’ recycled subduction fluids. The latter is supported by previous
strong evidence of mantle-derived N2/He contributions across the volcanic front of Nicaragua and Costa Rica.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction Caribbean, Cocos, Nazca, and South American (Sak et al., 2009). Between
the Cocos and Caribbean plates, convergence rate increases from
The Central America subduction zone is a complex deformation re- 60 mm/year (NW) to − 90 mm/year (SE) (DeMets, 2001; Peacock et
gion characterized by a rapid (70–90 mm/year) convergence rate of al., 2005; Hoernle et al., 2008; Protti et al., 2012). Through this margin,
young (15–25 Ma) oceanic lithosphere (DeMets, 2001; Peacock et al., deep earthquakes record the rupture at the interface of the Cocos-Carib-
2005; Hoernle et al., 2008), whereby the convergence rate and age of bean plates, defining a zone that marks the subducting Cocos plate up to
the incoming lithosphere vary slightly along-strike (Protti et al., 1995; 250 km beneath Nicaragua and about 80 km beneath the Costa Rican
DeMets, 2001), and mainly responds to the interaction of four plates: Volcanic front (Protti et al., 1995; Husen et al., 2003; Hoernle et al.,
2008; Arroyo et al., 2009).
⁎ Corresponding author at: Stable Isotope Research Group, Chemistry School,
Currently, the Costa Rica-Nicaragua subduction zone is a non-accret-
Universidad Nacional, Campus Omar Dengo, P.O. Box: 86-3000, Heredia, Costa Rica. ing margin, which is described by subduction erosion (von Huene and
E-mail address: ricardo.sanchez.murillo@una.cr (R. Sánchez-Murillo). Scholl, 1991; Ranero and von Huene, 2000; von Huene et al., 2000).

http://dx.doi.org/10.1016/j.jvolgeores.2017.08.013
0377-0273/© 2017 Elsevier B.V. All rights reserved.
A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183 173

The modern arc develops on a base of volcanic strata, which is related to (Alvarado, 2009), extending from NW with Orosí volcano (OR) to SE
arc magmatism in Nicaragua and the Caribbean Oceanic Plateau in Costa with Turrialba volcano (TU) (Fig. 1). The volcanic front of Costa Rica is
Rica (Denyer and Gazel, 2009 and references therein) The maximum el- divided into two main cordilleras (i.e. mountain ranges): Guanacaste
evation of volcanoes along Central America decreases from Guatemala Volcanic Cordillera (GVC), which is a NW-trending chain of volcanoes
(Tajamulco volcano: 4220 m a.s.l.) to Nicaragua (San Cristóbal volcano: (Orosí, OR; Rincón de la Vieja, RV; Miravalles, MV; and Tenorio, TE;
1745 m a.s.l.) with a decrease in crustal thickness (50 km to b 35 km) Alvarado, 2009) and the Central Volcanic Cordillera (CVC) (Platanar,
and variations in the basement geology from continental to oceanic PA; Poás, PO; Barva, BA; Irazú, IR; and Turrialba, TU). Between GVC
dominated characteristics, whereas in Costa Rica, volcanoes elevation and CVC, two twin isolated volcanoes exist: Arenal (AR) and Chato
increases (Irazú volcano: 3432 m a.s.l.) (45 km) near the amagmatic (not shown in Fig. 1, it is located next to AR) (Carr and Stoiber, 1990;
gap, then decreases (25 km) beneath central Panamá (La Yeguada vol- Herrstrom et al., 1995; Husen et al., 2003; Alvarado, 2009) (Fig. 1).
cano: 1297 m a.s.l.) (Leeman et al., 1994). The composition of the The GVC is considered a geothermal province with particular geo-
lavas in Nicaragua are typical of arc volcanoes (Feigenson and Carr, chemical signatures (Giggenbach and Corrales, 1992; Molina and
1993; Saginor et al., 2011; Saginor et al., 2013; Gazel et al., 2015), but Martí, 2016), which are transitional between an isotopically depleted
the volcanic front lavas in Costa Rica are well-known for their enriched mantle source and a high subduction signal characteristic of the Nicara-
ocean island basalt - like (OIB-like) signature (e.g., Reagan and Gill, guan volcanoes as well as the enriched mantle source and limited sub-
1989; Feigenson et al., 2004; Gazel et al., 2009). Recent studies have duction signal characteristic of the central Costa Rican volcanoes (Carr
provided convincing geochemical (i.e. rare trace elements) evidence and Stoiber, 1990; Herrstrom et al., 1995; Gazel et al., 2015). Multiple
that this anomalous OIB-like signature in the volcanic front is derived hydrothermal systems have been identified at the slopes (mainly Pacific
from the interaction of the mantle wedge with the Galapagos Hotspot facing slopes) of AR, RV, and MV volcanoes. In addition, PO, IR, and TU
tracks subducting beneath Costa Rica and Panama (Hoernle et al., volcanoes have hydrothermal systems with fluid discharge temperature
2008; Gazel et al., 2011, 2015). close to 90 °C (Giggenbach and Corrales, 1992; Fehn et al., 2002; Marini
This geochemical scenario in Nicaragua and Costa Rica offers the nec- et al., 2003; Zimmer et al., 2004; Moya et al., 2005; López et al., 2006;
essary heat and decompression cooling for storing fluids (i.e., meteoric Phillips-Lander et al., 2014).
water, sea water, magmatic water or mixed fluids including gas vents; Nicaragua (130,700 km2) is located between the Caribbean plate to
Molina and Martí, 2016) within fractured volcanic aquifers, which ulti- the east, and the Cocos plate to the west (Arengi and Hodgson, 2000).
mately constitute the main energy transport mechanism for surface hy- The active volcanic front (developed 350 ka ago; Carr et al., 2007;
drothermal manifestations. This hydrothermal upwelling provides heat, Brasse et al., 2015) is located in western Nicaragua (within the Cordille-
dissolved gases, and minerals to maintain a broad spectrum of ecosys- ra Los Marrabios), and it is represented by a great variety of volcanic for-
tems across the Pacific and Caribbean slopes of Costa Rica and Nicaragua, mations (i.e., active volcanoes, explosion craters, and scoria cones),
but also sustain a large set of hot-spring tourist activities and moderate along which most of the areas of geothermal priority are also located
hydrothermal energy production (Martínez-Tiffer et al., 1988; Balcazar (Martínez-Tiffer et al., 1988). The volcanism in Nicaragua is located
et al., 1993; Fridleifsson, 2001; Manzo, 2005). Although, stable isotope mainly in the Pacific slope of the country (Carr et al., 2004) (Cosigüina,
studies have been used to investigate hydrothermal manifestations in CS; San Cristóbal, SC; Telica, TE; El Hoyo, EH; Momotombo, MT;
Central America over three decades ago (Fournier et al., 1982; Goff et Apoyeque, AP; Masaya, MA; Mombacho, MO; Zapatera, ZP; Ciguatepe,
al., 1987; Giggenbach and Corrales, 1992; IAEA, 1992; Rowe et al., CI; Las Lajas, LA; Concepción, CO; and Maderas, MD; Fig. 1), and it is di-
1995; Nieva et al., 1997; Tassi et al., 2005; López et al., 2006; Birkle and vided into segments that are between 100 km and 300 km long (Carr et
Bundschuh, 2007; Rouwet et al., 2009; Molina and Martí, 2016), these ef- al., 2007; Saginor et al., 2011). Nicaragua has a region of active subsi-
forts were mainly focused on a specific hydrothermal field or volcano ed- dence, which is located roughly parallel to and within the Nicaraguan
ifice. To our knowledge, no spatial analysis has been undertaken to Depression (i.e., Lake Nicaragua) (Fig. 1), where sediments cover the
compare the contribution of recycled subduction fluids within the volca- bases of the active volcanoes (Saginor, 2008).
nic front of Central America using water stable isotopes and second order The Nicaraguan territory includes at least ten stratovolcanoes (in-
variables in hydrothermal system discharges. cluding nine active volcanoes, calderas, and pyroclastic cones; Simkin
In this study, stable isotope archives in hydrothermal waters of Costa et al., 2011). Conduction to the surrounding ground and the condensa-
Rica and Nicaragua were combined with new stable isotope data to a) tion of the gas are examples of losing heat mechanisms within hydro-
characterize the isotopic composition (δ18O, δ2H, d-excess, and lc-ex- thermal fluid manifestations, an example of the thermal structure is
cess), b) analyze the influence of secondary kinetic fractionation and me- the Masaya volcano, where diurnal temperature variations are b 5 °C,
teoric water inputs in the isotopic composition of hydrothermal waters, mainly due to the influence of rain events or variations about 10 °C relat-
and c) estimate the ‘andesitic water’ contribution (i.e. recycled subduc- ed to volcanic events (Pearson et al., 2008).
tion fluids) (Taran et al., 1989; Giggenbach, 1992) within the volcanic
front of Nicaragua and Costa Rica. This information provides new insights 2.2. Governing climatic features across Costa Rica and Nicaragua
regarding the use of stable isotopes and second-order variables as indica-
tors of hydrothermal/volcanic activity, as well as offering spatially-ori- Four regional air circulation patterns predominantly control the cli-
ented and quantitative evidence of subduction recycled fluid dynamics mate of Costa Rica: NE trade winds, the latitudinal migration of the Inter-
within the volcanic front of Central America. tropical Convergence Zone (ITCZ), cold continental outbreaks, and
sporadic influence of Caribbean cyclones (Waylen et al., 1996; Sáenz
2. Study area description and Durán, 2015). These circulation processes produce two predominant
rainfall maxima, one in May and June and the second in September–Oc-
2.1. Hydrothermal/volcanic activity across Costa Rica and Nicaragua tober, which are interrupted by a relative minimum in July–August
known as the Mid-Summer Drought (MSD) (Magaña et al., 1999;
Costa Rica (52,100 km2) is located in the southern extreme of Cen- Maldonado et al., 2013). In addition to these circulation processes, the
tral America to the west of the Caribbean plate, between the eastern Pa- continental divide of Costa Rica (i.e. a mountainous range that extends
cific Ocean and the Caribbean Sea. There is a convergent boundary, from NW to SE with a maximum elevation at the Chirripó peak:
where the Cocos plate is subducted beneath the Caribbean plate, 3820 m a.s.l.) also influences rainfall patterns across the country, dividing
which generated the volcanic front of Costa Rica (Fig. 1). In Costa Rica, the territory into the Caribbean and Pacific slopes. In general, annual
there are about fifteen dormant and five historically active volcanoes rainfall in Costa Rica varies from b 1500 mm in the drier northwestern re-
since 1700 to present, most of them are considered of a composite origin gion, 2500 mm in the Central Valley, and up to 7000 mm on the
174 A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183

Fig. 1. Overview of Central America showing the volcanic front of Nicaragua and Costa Rica. Volcanoes are classified according to the Global Volcanism Program, (2013) code (http://
volcano.si.edu/gvp_votw.cfm) from the Smithsonian Institution. Major subduction zones (solid-triangle line) and continental transform faults (dashed line) are also shown for
reference. Estimated convergence rate between Cocos and Caribbean plates taken from DeMets (2001) Peacock et al. (2005). Main volcanoes in Costa Rica: Orosí-Cacao (OR),
Miravalles (MI), Tenorio (TE), Arenal (AR), Platanar (PA), Poás (PO), Barva (BA), Irazú (IR), and Turrialba (TU). Main volcanoes in Nicaragua: Cosigüina (CS), San Cristóbal (SC), Telica
(TE), El Hoyo (EH), Momotombo (MT), Apoyeque (AP), Masaya (MA), Mombacho (MO), Zapatera (ZP), Concepción (CO), and Maderas (MD). Elevation gradient (m a.s.l.) is color coded.

Caribbean side of the Talamanca range (Fig. 2A). Temperature seasonal- excess; Landwehr and Coplen, 2006) required both isotopic values.
ity is low throughout the country (b2 °C). The mean annual temperature Costa Rica's archive dataset is composed of five sampling campaigns be-
on the coastal lowlands is about 27 °C, 20 °C in the Central Valley at tween 1982 and 2005, whereas Nicaragua's archive dataset comprises
around 1100 m a.s.l., and below 10 °C at the summits of the Talamanca seven sampling campaigns between 1968 and 2005. Historical stable iso-
range (Fig. 2B) tope analyses were conducted either by the International Atomic Energy
Similarly, Nicaragua's landscape is divided by the central highlands Agency (IAEA) or IAEA-approved laboratories using Isotope Ratio Mass
into the Pacific domain and the large extension of the Caribbean low- Spectrometry.
lands (Solé et al., 2016). Overall, the climate of Nicaragua is controlled Isotope records in rainfall were obtained from the Global Network of
by 1) macro-scale systems (i.e., the north-American continental anticy- Isotopes in Precipitation (GNIP, 2017). Although Costa Rica has an active
clone, the Azores's oceanic anticyclones, tropical cyclones, ITCZ, and rainfall monitoring network, only GNIP records were used to compare
ENSO), 2) meso-scale systems such as tropical waves, convective cells, the Local Meteoric Water Lines (LMWLs) of both countries. GNIP
and troughs; and 3) local systems, marine breezes, and mountain monthly records for Costa Rica (N = 679) are composed of 46 monitor-
waves (INETER, 2017). Due to its proximity to the Pacific Ocean, the rain- ing stations, ranging from 0 to 3000 m a.s.l. A detailed description of the
fall regime is strongly affected by ENSO events, resulting in large dry GNIP records of Costa Rica and sampling sites are presented in Sánchez-
spells across the Pacific domain, northern and northeastern central re- Murillo et al. (2013). GNIP monthly records from Nicaragua (N = 51; in-
gions. In general, annual rainfall in Nicaragua varies from b 800 mm in termittent monthly sampling between 1968 and 2005) are composed of
the drier northern and central regions, 2000 mm in the Pacific domain, 28 monitoring stations mainly on the Pacific domain and central high-
2500 mm on the windward slope of the central highlands, and up to lands, ranging from 22 to 1300 m a.s.l. The remoteness and difficult ac-
5000 mm on the Caribbean side (Fig. 2A). Mean annual temperature in cess has historically limited isotope observations across the Caribbean
Nicaragua is about 25.4 °C. High dry season temperatures could reach lowlands of Nicaragua.
up to 40 °C in the central and northern regions (INETER, 2017) (Fig. 2B). In addition, raster grids of mean annual δ18O composition for Nicara-
gua and Costa Rica were used (ArcGIS 10.4, ESRI, USA) to compare spatial
3. Stable isotope data compilation and analysis trends over the volcanic front of both countries; insufficient historical
and recent isotopic data in Nicaragua limit the computation of rainfall
Stable isotope archives of hydrothermal fluids of Costa Rica (N = isoscapes with a greater resolution. Raster grids were extracted and
340) and Nicaragua (N = 309) were obtained directly from the Water adapted from Bowen and Revenaugh (2003):
Resources Programme of the International Atomic Energy Agency http://wateriso.utah.edu/waterisotopes/pages/data_access/ArcGrids.
(IAEA) (Fig. 3, see Supplemental material database S1). Site characteris- html
tics within IAEA archives are limited to the historical information avail- In addition, 70 recent samples of hydrothermal waters (mostly chlo-
able. Sampling sites comprised few mud-pools such as the ‘San Jacinto ride and acid-sulfate brines) (collected and analyzed between 2015 and
Hervideros’ within the Telica volcano (Nicaragua) and ‘Pailas at the 2016) from Costa Rica were included in the database (Fig. 3; see Supple-
Rincón de la Vieja volcano (Costa Rica), hot springs (over 70% of the sam- mental material database S1). Samples were collected as close as possible
ples), and few drilled and dug wells. Samples were collected following to the most prominent discharge point, avoiding pools with potential
IAEA sampling procedures (available at: http://www-naweb.iaea.org/ steady-state evaporation. Samples were filtered using a Midisart PTFE
na/index.html). Only samples containing both δ18O and δ2H were used, (polytetrafluorethylene) 0.45 μm syringe membrane (Sartorius AG, Ger-
since second-order calculations such as deuterium excess (hereafter, d- many), transferred to 30 mL HDPE (high-density polyethylene) vials,
excess; Dansgaard, 1964) and the line-conditioned excess (hereafter lc- and stored at 5 °C until analysis. Stable isotope analysis was conducted
A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183 175

Fig. 2. A) Mean annual rainfall (MAR, color coded in mm/yr) for Costa Rica (Sánchez-Murillo and Birkel, 2016) and Nicaragua (INETER, 2005). B) Mean annual temperature (MAT, °C) for
Costa Rica (adapted from Esquivel et al., 2017) and Nicaragua (INETER, 2005).

at the Stable Isotope Research Group facilities of the Universidad Nacional = 7.43, b = 5.2). This calculation uses the LMWLs as a reference rather
(Heredia, Costa Rica) using a Cavity Ring Down Spectroscopy water iso- than simply using the Euclidian distance from the Global Meteoric
tope analyzer L2120-i (Picarro, USA). The secondary standards were: Water Line (GMWL; Craig, 1961) (Sprenger et al., 2017). Negative values
Moscow Tap Water, MTW (δ2H = −131.4‰, δ18O = −17.0‰), Deep of lc-excess indicate kinetic fractionation and are often found in isotopi-
Ocean Water, DOW (δ2H = −1.7‰, δ18O = −0.2‰), and Commercial cally-enriched waters. The coefficients a and b are the slope and the y-in-
Bottled Water, CAS (δ2H = −64.3‰, δ18O = −8.3‰). MTW and DOW tercept of the LMWLs, respectively. The y-intercept (b) and slope (a)
standards were used to normalize the results to the VSMOW2-SLAP2 were obtained from Costa Rica (Sánchez-Murillo et al., 2013) and Nicara-
scale, while CAS was used as a quality control and drift control standard. gua isotope records from the GNIP database (GNIP, 2017). Based on the
The analytical long-term uncertainty was: ±0.5 (‰) (1σ) for δ2H, ±0.1 analytical precision reported, the estimated average uncertainties are
(‰) (1σ) for δ18O. Stable isotope compositions are presented in delta no- ±1.1‰ (d-excess) and ±1.5‰ (lc-excess).
tation δ (‰, per mil), relating the ratios (R) of 18O/16O and 2H/1H, relative
to Vienna Standard Mean Ocean Water (V-SMOW).
The d-excess was calculated for each sample following Eq. (1). In ad- d−excess ¼ δ2 H−8∙δ18 O ð1Þ
dition, to determine the degree of deviation of hydrothermal waters
from the LMWLs, lc-excess was calculated according to Landwehr and
Coplen (2006) (Eq. (2)) (Costa Rica: a = 7.61, b = 7.4; Nicaragua: a lc−excess ¼ δ2 H−a∙δ18 O−b ð2Þ
176 A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183

Fig. 3. Distribution of sampling sites within the volcanic front of Nicaragua (N = 309) and Costa Rica (N = 419) (green dots).

Following Giggenbach (1992) and Giggenbach and Corrales (1992), but it was only used to represent the ‘andesitic water’ space in the dual
the fraction of ‘andesitic water’ (Xa) (i.e., waters associated with more si- isotope plot. A Kruskal-Wallis non-parametric one way analysis of vari-
licic andesitic magmatism typical of mature volcanic arc systems; Taran ance on ranks (Kruskal and Wallis, 1952) was applied to test if there
et al., 1989) was calculated for each hydrothermal sample exhibiting was stochastic dominance regarding the ‘andesitic contribution’ values
lc-excess values b 0 (i.e., indicator of kinetic fractionation; see Supple- between Nicaragua and Costa Rica. A significant difference was deter-
mental material database S2), using an end-member mixing model pre- mined (p b 0.001) when the median values were greater than expected
sented in Eq. (3): by chance. In addition, a multiple comparison procedure was applied
using Dunn's method (Dunn, 1961) to test if there is evidence of stochas-
ðδh −δm Þ tic dominance between the samples. Dunn's method approximates exact
Xa ¼ ð3Þ
ðδa −δm Þ rank-sum test statistics by using the mean rankings of the results in each
group from the previous Kruskal-Wallis non-parametric test and pro-
where δh is the δ18O (‰) of any hydrothermal water, δm corresponds to vides an inference in mean ranks in each group. Density distribution
the mean annual δ18O (‰) in meteoric water (Table 1), and δa denotes analysis between the stale isotope and second-order variables was con-
the widely-used (e.g. Hedenquist and Lowenstern, 1994; Chiodini et al., ducted using the statistical programming language R (R Core Team,
2001; Tassi et al., 2005; Birkle and Bundschuh, 2007) δ18O (‰) reference 2014).
composition for ‘andesitic water’ isotope space: + 10 ± 3‰
(Giggenbach, 1992). Due to the high collinearity between δ18O and δ2H 4. Results
(r = 0.99), only δ18O values were used for end-member calculations.
The expanded uncertainty for Xa corresponded to ±4.8%. The reference 4.1. Stable isotope characteristics in meteoric waters
value of δ2H in ‘andesitic water’ is − 20 ± 10‰ (Giggenbach, 1992),
Mean annual δ18O in rainfall of Costa Rica ranged from −10.1‰ to
−1.8‰ with an overall mean of −6.7 ± 1.4‰, while δ2H ranged from
Table 1
Summary of isotopic composition (‰) and second-order variables (‰) in rainfall and hy- − 17.3‰ to − 70.4‰ with a mean of − 43.2 ± 10.3‰. The d-excess
drothermal/volcanic systems of Costa Rica and Nicaragua. values varied from +2.9‰ to +14.4‰ with a mean of +10.8‰. In gen-
eral, the meteoric water line of Costa Rica is described as: δ2H =
Costa Rica
7.61·δ18O + 7.40 (r2 = 0.98, p b 0.001, N = 679; Sánchez-Murillo et
Variables (‰) Hydrothermal water Rainfall al., 2013) (Fig. 4A; Table 1). In Costa Rica, a clear isotopic separation in
Mean ± σ Max Min Mean ± 1σ Max Min rainfall ratios is observed across the Pacific and Caribbean domains,
δ18O −4.5 ± 2.4 +6.2 −11.5 −6.7 ± 1.4 −1.8 −10.1 which is explained by the prevailing trade winds direction (NE → SW),
δ2H −34.2 ± 9.0 +7.6 −82.3 −43.2 ± 10.3 −17.3 −70.4 two distinct moisture sources (Caribbean Sea and eastern Pacific
d-excess +1.8 ± 9.8 +28.4 −63.1 +10.8 ± 2.0 +14.4 +2.9 Ocean), distinct rainfall regimes (spatially and temporally), and the oro-
lc-excess −7.3 ± 9.0 +18.4 −68.2 +0.4 ± 2.5 +4.8 −10.6 graphic distillation across the central mountain range (Sánchez-Murillo
Nicaragua
et al., 2016) (Fig. 4D).
Mean annual δ18O in rainfall of Nicaragua ranged from −10.2‰ to
Variables (‰) Hydrothermal water Rainfall
− 2.9‰ with a mean of − 6.4 ± 1.9‰, while δ2H ranged from
Mean ± SD Max Min Mean ± SD Max Min −72.5‰ to −12.2‰ with a mean of −42.2 ± 14.1‰. Deuterium excess
δ18O −5.2 ± 2.1 +5.8 −11.2 −6.4 ± 1.9 −2.9 −10.2 values varied from +3.3‰ to +18.7‰ with a mean of +8.9‰ (Table 1).
δ2H −45.6 ± 7.8 −8.2 −66.4 −42.2 ± 14.1 −12.2 −72.5 The meteoric water line of Nicaragua is described as: δ2H = 7.43·δ18O
d-excess −4.1 ± 12.3 +28.7 −54.8 +8.9 ± 3.3 +18.7 +3.3 + 5.2 (r2 = 0.95, p b 0.001, N = 52) (Fig. 4A; Table 1). Regarding the
lc-excess −12.3 ± 11.2 +17.2 −56.6 −0.02 ± 3.17 +7.8 −5.7
moisture transport processes, Central America is affected by the direct
A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183 177

Fig. 4. A) Dual stable isotope plot for hydrothermal systems in Nicaragua (red down-triangles) and Costa Rica (blue up-triangles). The GMWL (black line) is shown as reference. Costa Rica
(blue dashed line) (Sánchez-Murillo et al., 2013) and Nicaragua (red dashed line) (GNIP, 2017) LMWLs are also included. The filled-rectangle denotes the end-member ‘andesitic water’
space described by Giggenbach (1992). B and C) Density distribution plots for δ18O (‰) and δ2H (‰) in hydrothermal systems of Costa Rica and Nicaragua; dashed lines represent the
median value. D) Mean annual δ18O spatial variability in rainfall adapted from Bowen and Revenaugh, 2003).

influence of vapor fluxes from the Caribbean Sea and the eastern tropi- 4.2. Isotopic deviation within the hydrothermal/volcanic systems
cal Pacific Ocean. The semi-closed basin of the Caribbean Sea (δ2H =
−75.0‰ and δ18O = −10.6‰, extracted from Good et al., 2015) has a In Costa Rica, δ18O in hydrothermal systems ranged from −11.5‰ to
more enriched marine isotope vapor ratio than the eastern tropical Pa- +6.2‰ with a mean of −4.5 ± 2.4‰, whereas δ2H composition varied
cific Ocean (δ2H = − 100.0‰ and δ18O = − 13.75‰, extracted from from − 82.3‰ to +7.6‰ with a mean of −34.2 ± 9.0‰ (Fig. 4 B and
Good et al., 2015) with an overall difference of − 25‰ (δ2H) and C; Table 1). The hydrothermal water line (HWL) (i.e., evaporation line)
− 3.2‰ (δ18O). Similarly, δ18O ratios in seawater across the Central of Costa Rica can be described as: δ2H = 4.70·δ18O − 13.0 (r2 = 0.79,
America region show a relative difference of 1–2‰ between the Carib- N = 41, p b 0.001) (Fig. 4A). Tassi et al. (2005), reported a HWL for the
bean Sea and the eastern tropical Pacific Ocean (LeGrande and Rincón de la Vieja (RV) hydrothermal field of δ2H = 2.70·δ18O + 12.2
Schmidt, 2006). (N = 16). Similarly, d-excess varied from −63.1‰ to +28.4‰ with a
These dissimilarities are well depicted in the rainfall isoscape of Costa mean of + 1.8 ± 9.8‰ and lc-excess values ranged from − 68.2‰ to
Rica and Nicaragua (Fig. 4D), where the most enriched values occurred in +18.4‰ with a mean of −7.3 ± 9.0‰ (Fig. 5 A, B, and C; Table 1). In
the Caribbean lowlands, while the most depleted values lie along the in- the case of Costa Rica, a clear bimodal distribution is observed among
terior mountain range and along the Pacific coast. The weakening of the the hydrothermal water samples. The latter is explained by the strong
trade winds from May to November allows the incursion of Pacific mois- isotopic separation in rainfall between the Pacific and Caribbean slopes
ture to the Central American Isthmus, resulting in intense convective (Fig. 5 D and E) and further emphasizes the importance of recharge
rainfall during May and June, and throughout August to October mixing from two distinct meteoric water inputs.
(Maldonado et al., 2013) with more depleted isotopic values. From De- For Nicaragua, δ18O composition ranged from − 11.2‰ to + 5.8‰
cember to April, the trade winds present their maximum velocities pro- with a mean of − 5.2 ± 2.1‰, while δ2H composition varied from
viding intense monthly rainfall and near-uniformly enriched isotope − 66.4‰ to − 8.2‰ with a mean of − 45.6 ± 7.8‰ (Fig. 4B and C,
values across the Caribbean lowlands. Although, the prevailing trade Table 1). The HWL of Nicaragua can be described as: δ2H = 2.71·δ18O
wind direction across Nicaragua is similar to Costa Rica, the absence of − 31.6 (r2 = 0.52, N = 309) (Fig. 4A). Elkins et al. (2006) reported a
a more continuous and high mountain range barrier, results in a more slope of 3.6 in Momotombo (MO) vapor samples and suggested a large
uniform isotopic composition in rainfall across Nicaragua. Enrichment degree of mixing between regional meteoric water and a magmatic
is observed from the Caribbean lowlands towards the Nicaraguan De- vapor component. In Nicaragua, d-excess composition ranged from
pression and depletion is observed at the windward slope of the volcanic −54.8‰ to +28.7‰ with a mean of −4.1 ± 12.3‰ and lc-excess varied
front and within the northern central region, close to the Honduran bor- from −56.6‰ to +17.2‰ with a mean of −12.3 ± 11.2‰ (Fig. 5 A, B,
der (Fig. 4D). and C). Contrary to the observed bimodal distribution of Costa Rica's
178 A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183

Fig. 5. A) Relationship between δ2H (‰) and d-excess (‰) in hydrothermal systems across Nicaragua and Costa Rica. The dashed line represents the global mean (+10‰) for d-excess. B)
Relationship between δ2H (‰) and lc-excess (‰) in hydrothermal systems across Nicaragua and Costa Rica. The dashed line represents lc-excess value of zero. The light-blue box contains
a group of samples sharing isotopic similarity with meteoric conditions, while the red box denotes samples with a large evaporative effect and interaction with recycled subduction fluids.
C) Box plots showing d-excess (‰) and lc-excess (‰) distribution for Nicaragua and Costa Rica including median, 25th and 75th percentiles, and error bars. Crosses denote the number of
outliers. D and E) Density distribution plots for d-excess (‰) and lc-excess (‰) in hydrothermal systems of Costa Rica and Nicaragua; dashed lines represent the median value.

hydrothermal waters, in Nicaragua this effect is slightly attenuated by b 0), with exceptions for crater water and mud-pool samples (Poás crater
the overall uniformity of the meteoric water input across the volcanic water, ‘San Jacinto Hervideros’ and ‘Pailas’), where steady-state evapora-
front (Fig. 5 D and E) due to the absence of a strong orographic effect. tion largely affects the isotopic deviation and resulted in greater appar-
ent ‘andesitic water’ contributions.
4.3. End-member mixing calculations
4.4. Isotopic spatial variability across the hydrothermal/volcanic systems
The combination of a magmatic end-member (Giggenbach, 1992),
well constrained rainfall isotopic ratios (isoscapes), and a large spatial In Costa Rica, more enriched δ18O and δ2H values occurred along the
coverage of hydrothermal samples, allowed the calculation of ‘andesitic GVC (northern Pacific region) (Fig. 1.), while the most depleted values
water’ (recycled subduction fluids) contribution within the volcanic were observed along the Pacific slope and CVC (Fig. 7A, showing only
front of Nicaragua and Costa Rica. In Costa Rica, the percentage of contri- δ18O, since high collinearity with δ2H reproduced the same spatial pat-
bution ranged from 0.12 (%) to 77.3 (%) with a mean of 19.7 ± 10.3 (%) tern). The latter partially relies on the isotopic distribution of rainfall in
and a median value of 22.3 (%). For Nicaragua, the estimated ‘andesitic the country, where depleted compositions are a normal feature in the Pa-
water’ input ranged from 0.6 (%) to 74.4 (%) with a mean of 15.3 ± cific domain and high elevation mountains (central and southeastern re-
10.8 (%) and a median value of 14.6 (%) (Fig. 6A). A Kruskal-Wallis gions) (Fig. 4D). Topographic depressions between volcano edifices or –
non-parametric analysis of variance followed by an all pairwise compar- in general – along the main mountain range (crossing the country in
ison determined the statistically significant difference (p b 0.001) be- NW-SE direction) facilitate the transport of air masses from the Caribbe-
tween the median values in both countries. Density distribution an domain, resulting in the contribution of enriched Caribbean-type rain-
analysis (Fig. 6B) exhibited a particular threshold of ‘andesitic water’ fall to the Pacific slope volcanic aquifers (Sánchez-Murillo et al., 2016)
contribution close to 40% (98th percentile of the samples with lc-excess and further enhanced a bimodal distribution within the hydrothermal
A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183 179

isotopic exchange of δ18O during rock-water interactions, consequently,


also showed the imprint water-rock ratios (Craig, 1963; Blattner, 1993).
These highly evolved waters (i.e. completely equilibrated fluids with
hosting rocks at high temperatures) often exhibit δ18O and δ2H so pro-
foundly modified, which in turn, can no longer be used to determine
the fluid origin (Criss, 1999). In addition, these early studies omitted
or considered the ‘andesitic water’ contribution to be insignificant
(b1%), therefore, unlikely to be responsible of the observed isotopic de-
viations from meteoric water conditions.
Taran et al. (1989) used the name ‘andesitic water’ to describe the
quite distinct isotopic composition of fumaroles condensates from volca-
noes in Kamchatka (Russia), and the authors associated HWLs with
slopes close to 2–3 to non-equilibrium evaporation (at near boiling tem-
peratures) from steam-heated pools. Giggenbach (1992) provided a
widely used (e.g. Hedenquist and Lowenstern, 1994; Chiodini et al.,
2001; Tassi et al., 2005; Birkle and Bundschuh, 2007) isotopic range for
‘andesitic water’ condensates, limiting the isotopic composition to −20
± 10‰ for δ2H and −10 ± 3‰ for δ18O (Fig. 4A) and reported mean ‘an-
desitic water’ contribution in the northern Pacific region of the volcanic
front of Costa Rica close to 18% (Giggenbach and Corrales, 1992). Within
the volcanic front of Nicaragua, the HWLs (δ2H = 2.71·δ18O − 31.6; r2 =
0.52, N = 309) (Fig. 4A), clearly indicated a notable departure from me-
Fig. 6. A) Relationship between ‘andesitic water’ contribution (%) and lc-excess (‰) in teoric conditions due to potential intense kinetic fractionation, which is
Costa Rica (blue up-triangles) and Nicaragua (red down-triangles). Grey-dashed line
reflected in the consistent and significant (p b 0.001) low d-excess and
represents the 98th percentile with a maximum value of 40%. B) Density distribution
plot for the ‘andesitic water’ contribution (%) in Costa Rica and Nicaragua; dashed lines lc-excess median values (Fig. 5D and E). An important aspect is that
represent the median value. low relative humidity (40–60%) and high dry season temperature (up
to 40 °C) within the Pacific slope of Nicaragua (i.e., part of the Dry Corri-
dor of Central America; Hidalgo et al., 2017) may also enhance surface ki-
systems that experienced rainfall regimes influenced by Pacific and Ca- netic fractionation.
ribbean moisture. A direct comparison of the slope and intercept ratios in the HWLs of
Deuterium excess and lc-excess exhibited less spatial variability with Nicaragua and Costa Rica suggested a greater kinetic fractionation that
low values (Fig. 7B and C) from active volcanoes such as Rincón de la coupled with drier/warmer near-surface ambient conditions ultimately
Vieja (RV, Las Pailas mud pool, d-excess = − 43.1‰, lc-excess = − enhances the deviation from local meteoric compositions in Nicaragua
48.1‰) and Poás (PO, crater water: d-excess = −63.1‰, lc-excess = (slopeNIC/slopeCR = 0.58; interceptNIC/intercepCR = 2.43). Nevertheless,
−68.2‰) (Fig. 8A and C). Overall, d-excess and lc-excess indicated the the end-member mixing model (Giggenbach, 1992) estimated slightly
predominance of two distinct types of hydrothermal fluids: a) isotopical- greater ‘andesitic water’ contribution in the hydrothermal systems of
ly-altered by kinetic fractionation (particularly within the GVC) and b) Costa Rica (mean = 19.7 ± 10.3 (%)) compared to Nicaragua (mean
isotopically-close to meteoric conditions showing d-excess values near = 15.3 ± 10.8 (%) (Fig. 6A and B). However, a Kruskal-Wallis test deter-
+10‰ or greater as a result of a common enhanced moisture recycling mined the statistically significant difference (p b 0.001) between the me-
and the wetter rainfall regime in the CVC and Pacific lowlands. dian values of the estimated ‘andesitic water’ contribution in both
Isotope compositions across the volcanic front of Nicaragua were fair- countries. Only few samples in both countries were above the density
ly variable and exhibited no particular isotopic trend when compared to distribution threshold of 40% (grey dashed line, Fig. 6A) in terms of ‘an-
the rainfall isoscape (Fig. 4D). Enriched δ18O values were concentrated in desitic water’ contribution, exclusively for water samples from large hy-
the ‘San Jacinto Hervideros’ (i.e., boiling mud pools; close to Telica volca- drothermal fields (boiling mud pools, Fig. 8 A, B, and C) and active craters
no, TE) and within the Momotombo volcano (MT) hydrothermal field (e.g. Poás volcano, PO), which also experienced steady-state evaporation
(δ18O ranging from +3.2 up to + 5.8‰) (Fig. 7D and 8C). Contrary to of brine fluids at elevated temperatures (Rowe et al., 1995; Rouwet et al.,
Costa Rica's hydrothermal fluids, where two distinct groups were identi- 2016).
fied, in Nicaragua 87% of the samples presented lc-excess and d-excess In general, the isotopic composition and second order variables
values below 0, suggesting strong presence of secondary evaporation presented a clear bimodal distribution in Costa Rica's hydrothermal
processes (Fig. 7E and F); however, the limited number of samples systems (Fig. 4 A and B, and 5 D and E) where a clear spatial pattern
from deep wells restricted the possibility of separating surface evapora- is evident, whereas in Nicaragua the distribution of isotopic compo-
tion from underground fractionation effects. However, only 8% of the sition is less influenced by the meteoric isotopic input but largely af-
samples exhibited d-excess values greater than +10‰, reinforcing the fected by kinetic fractionation and potentially near-surface ambient
evidence of the drier conditions within the Dry Corridor of Nicaragua conditions. In addition, relative poor correlations (Costa Rica, r2 =
(Hidalgo et al., 2017) and the limited subsidence of recycled moisture. 0.22; Nicaragua, r2 = 0.31) between δ18O and the elevation gradient
Climate conditions along the Pacific slope of Nicaragua represented by suggested that contribution of ‘andesitic water’ and subsequent ki-
a drier rainfall regime (i.e., less potential meteoric water mixing) and netic fractionation are the main drivers of the observed isotopic de-
greater ambient temperatures (Fig. 2A and B), increase the potential of parture from the expected orographic distillation trend captured in
surface kinetic fractionation of the hydrothermal fluids. the rainfall isoscapes.

5. Discussion 5.2. Meteoric and recycled subduction fluids: groundwater perspective


within the hydrothermal/volcanic systems
5.1. Isotopic deviation and recycled subduction fluids
The inherent complexity of volcanic-originated aquifers, particularly,
Early descriptions of isotopic deviations (between 3 and 5‰ in δ18O) in steep and highly fractured groundwater reservoirs, whereby lateral
from LMWLs in high temperature waters were attributed mainly to and vertical meteoric water mixing occurs (Madrigal-Solís et al., 2017),
180 A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183

Fig. 7. Isoscapes of hydrothermal waters of Nicaragua and Costa Rica: A–B) Spatial variability of δ18O (‰). C–D) Spatial variability of d-excess (‰). E–F) Spatial variability of lc-excess (‰).
Volcanoes acronyms are described in Fig. 1 and Section 2.1. Elevation gradient (m a.s.l.) is color coded.

challenges existing models of subsurface water flow paths and storage water contribution of 83.6% within this geothermal reservoir. In the
(Delcamp et al., 2016) in such systems. In terms of the ‘andesitic water’ case of Costa Rica, the meteoric water contribution to groundwater sys-
interaction with meteoric waters, stable isotopes revealed a maximum tems – at moderate and high altitudes within the volcanic front – is com-
contribution threshold of 40% (98th percentile) (Fig. 6A). Elkins et al. posed of two rainfall types: depleted Pacific-type (mean annual value
(2006), reported mantle signature contribution to Nicaraguan volcanic across the Pacific domain = − 7.0 ± 1.0‰) and enriched Caribbean-
and geothermal gases of 54%, based on N2/He measurements at the type (mean annual value across the Caribbean domain = − 5.0 ±
Mombacho volcano (MO, Nicaragua). Giggenbach and Corrales (1992) 2.4‰) (Sánchez-Murillo et al., 2016). The wetter rainfall regime across
reported up to 18% of ‘andesitic water’ contribution to hydrothermal the Costa Rica's Pacific slope (roughly 1000 mm greater than Nicaragua's
fluids within the MV, TE, and RV volcanoes in the northern Pacific por- Pacific domain, produces a masking effect reflected in a group of samples
tion of the volcanic front of Costa Rica. More recently, Molina and Martí plotted along the LMWLs with a regression line of δ2H = 7.97·δ18O
(2016) estimated a magmatic water contribution in the western slope + 12.2; r2 = 0.99, N = 95, 22.6% of total samples), whereas such
of RV volcano of approximately 16.4%, which resulted in a meteoric masking cannot be identified in Nicaragua.
A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183 181

Fig. 8. Photographs of hydrothermal systems from Costa Rica and Nicaragua: A) Las Pailas mud pool, Rincón de la Vieja (RV) volcano, Costa Rica (Courtesy of M. Martínez and Wendy Saéz,
OVSICORI, Universidad Nacional, Costa Rica). B) Active degassing at Las Pailas mud pool, Rincón de la Vieja (RV) volcano, Costa Rica (Courtesy of G. Alvarado-Induni, ICE, Costa Rica). C) San
Jacinto ‘Hervideros’ (mud boiling pool), Telica (TE) volcano hydrothermal field, Nicaragua (source: www.vianica.com). D) Jaulares hydrothermal spring, northeastern flank of the Irazú
(IR) volcano, Costa Rica. The yellow-squared inset shows a zoom-in of iron oxides deposits at 36 °C (Courtesy of R. Sánchez-Murillo, Stable Isotope Research Group, Universidad
Nacional, Costa Rica).

5.3. Recycled subduction fluids origin respectively. Further validation of this stable isotope approach should in-
volve weekly to monthly sampling in hydrothermal fields including sur-
The source of the ‘andesitic water’ is most likely recycled seawater, face water discharge and deep/shallow wells to better constrain the
which enters the subduction Central America zone as water bound in hy- temporal evolution of the isotopic ratios (both in rainfall and hydrother-
drous mineral subducting slab (Giggenbach, 1992; Ryan and Chauvel, mal inputs) during the rainy and dry seasons, but most importantly, to
2013; Zamboni et al., 2016). The large spectrum of ‘andesitic water’ con- separate the effect of surface kinetic fractionation versus underground
tribution (Fig. 6A) within the hydrothermal systems of Costa Rica and evaporation and mixing with ‘pre-shifted’ recycled subduction fluids.
Nicaragua largely reflects the degree of interaction and exchange with Lastly, the spatial and isotope-informed indicators presented may be
isotopically ‘pre-shifted’ recycled fluids (i.e. fluids generally higher in used to better understand the hydrothermal/volcanic activity as well as
δ18O than the LMWLs). The latter is supported by several studies of providing new spatially-oriented insights of recycled fluids from the
helium ratios (3He/4He) across the volcanic front of Costa Rica and Nica- Central America subduction zone.
ragua (Fischer et al., 2002; Shaw et al., 2003; Elkins et al., 2006; Lucic et Supplementary data to this article can be found online at http://dx.
al., 2014; Rizzo et al., 2016), which provided strong evidence of upper doi.org/10.1016/j.jvolgeores.2017.08.013.
mantle (8 ± 1 RM/RA; RM = ratio in the sample and RA ratio of air) contri-
butions to the Central America volcanism. Our results suggested that
these upper mantle wedge inputs covered a maximum contribution of Acknowledgements
40% in the sloping (i.e. aquifers within the inter-mountainous topo-
graphic slopes) volcanic aquifers of Costa Rica and Nicaragua. This project was supported by International Atomic Energy Agency
grants COS7005 (Ensuring Water Security and Sustainability of the Cen-
tral Valley of Costa Rica) and CRP-19747 to RSM under the initiative “Sta-
6. Conclusions
ble isotopes in precipitation and paleoclimatic archives in tropical areas
to improve regional hydrological and climatic impact models” and the
Hydrothermal systems of Nicaragua exhibited a notable departure
Research Office of the Universidad Nacional of Costa Rica through grants
from meteoric conditions due to potential intense kinetic fractionation,
SIA-0482-13, SIA-0378-14, and SIA-0101-14. The authors would also like
which is reflected in the consistent low d-excess and lc-excess values.
to thank the IsoNET initiative funded by the University of Costa Rica Re-
In Costa Rica, hydrothermal fluids are divided in two main groups: a) iso-
search Council and the recent sampling help from OVSICORI
topically-close to meteoric conditions and b) isotopically-altered by the
(Universidad Nacional, Heredia, Costa Rica) and Instituto Costarricense
interaction with recycled subduction fluids. In fact, using the threshold
de Electricidad (San José, Costa Rica). We also thank the Programme on
of lc-excess equal 0, two clearly distinct regression lines can be distin-
Water Resources of the International Atomic Energy Agency for granting
guished for Costa Rica: a) δ2H = 7.97·δ18O + 12.2, which correspond
access to stable isotope archives of Costa Rica and Nicaragua. A complete
to meteoric conditions (Sánchez-Murillo et al., 2013) and b) δ2H =
database is provided in the supplemental materials.
4.65·δ18O − 15.0, with lc-excess values ranging from − 0.3‰ up to
−68.2‰ (e.g. Poás volcano crater). Greater rainfall inputs in Costa Rica
References
with respect to Nicaragua resulted in the partial attenuation of the evap-
orative effect, particularly within the Pacific slope. End-member mixing Alvarado, G.E., 2009. Los volcanes de Costa Rica: geología, historia, riqueza natural y su
calculations indicated a significant (p b 0.001) difference between the gente. [Volcanoes of Costa Rica: geology, history, natural diversity, and people].
EUNED, San José, Costa Rica, p. 355.
mean ‘andesitic water’ contribution within the hydrothermal systems Arengi, J.T., Hodgson, G.V., 2000. Overview of the geology and mineral industry of Nicara-
of 15.3 ± 10.8 (%) and 19.7 ± 10.3 (%) for Nicaragua and Costa Rica, gua. Int. Geol. Rev. 42 (1):45–63. http://dx.doi.org/10.1080/00206810009465069.
182 A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183

Arroyo, I.G., Husen, S., Flueh, E.R., Gossler, J., Kissling, E., Alvarado, G.E., 2009. Three-dimen- Giggenbach, W.F., Corrales, R., 1992. Isotopic and chemical composition of water and
sional P-wave velocity structure on the shallow part of the Central Costa Rican Pacific steam discharges from volcanic-magmatic-hydrothermal systems of the Guanacaste
margin from local earthquake tomography using off-and onshore networks. Geophys. Geothermal Province, Costa Rica. Geochemistry 7 (4):328–329. http://dx.doi.org/
J. Int. 179 (2):827–849. http://dx.doi.org/10.1111/j.1365-246X.2009.04342.x. 10.1016/0883-2927(92)90022-U.
Balcazar, M., Gonzalez, E., Ortega, M., Flores, J.H., 1993. Geothermal energy prospecting in Global Volcanism Program, 2013. In: Venzke, E. (Ed.), Volcanoes of the World, v. 4.6.0.
El Salvador. Nucl. Tracks Radiat. Meas. 22 (1–4):273–276. http://dx.doi.org/10.1016/ Smithsonian Institution Downloaded 16 Jun 2017. http://dx.doi.org/10.5479/
0969-8078(93)90065-C. si.GVP.VOTW4-2013 Accessed on June 1, 2017. Available at. http://volcano.si.edu/
Birkle, P., Bundschuh, J., 2007. Hydrogeochemical and isotopic composition of geothermal gvp_votw.cfm.
fluids. Central America: Geology, Resources and Hazards. Taylor & Francis, pp. 780–838. GNIP/International Atomic Energy Agency, 2017. Global network of isotopes in precipita-
Blattner, P., 1993. “Andesitic water”: a phantom of the isotopic evolution of water-silicate tion and global network of isotopes in river, The GNIP Databases. [Available at http://
systems. Comment on “isotopic shifts in waters from geothermal and volcanic systems www.iaea.org/water, accessed on 21, April, 2017].
along convergent plate boundaries and their origin” by WF Giggenbach. Earth Planet. Goff, F.E., Truesdell, A.H., Grigsby, C.O., Janik, C.J., Shevenell, L.A., Paredes, J.R., Gutierrez,
Sci. Lett. 120 (3–4):511–518. http://dx.doi.org/10.1016/0012-821X(93)90261-7. J.W., Trujillo, J., Counce, D.A., 1987. Hydrogeochemical Investigation of Six Geother-
Bowen, G.J., Revenaugh, J., 2003. Interpolating the isotopic composition of modern meteoric mal Sites in Honduras, Central America (No. LA-10785-MS). Los Alamos National
precipitation. Water Resour. Res. 39 (10). http://dx.doi.org/10.1029/2003WR002086. Lab., NM (USA) http://dx.doi.org/10.2172/6268112.
Brasse, H., Schäfer, A., Díaz, D., Alvarado, G.E., Muñoz, A., Mütschard, L., 2015. Deep-crustal Good, S.P., Noone, D., Kurita, N., Benetti, M., Bowen, G.J., 2015. D/H isotope ratios in the
magma reservoirs beneath the Nicaraguan volcanic arc, revealed by 2-D and semi 3-D global hydrologic cycle. Geophys. Res. Lett. 42:5042–5050. http://dx.doi.org/10.1002/
inversion of magnetotelluric data. Phys. Earth Planet. Inter. 248:55–62. http:// 2015GL064117.
dx.doi.org/10.1016/j.pepi.2015.08.004. Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation of hydro-
Carr, M.J., Stoiber, R.E., 1990. Volcanism. In: Dengo, G., Case, J. (Eds.), The Geology of North thermal ore deposits. Nature 370 (6490), 519–527.
America. Geol. Soc. of Am., pp. 375–391. Herrstrom, E.A., Reagan, M.K., Morris, J.D., 1995. Variations in lava composition asso-
Walker, J.A., 2004. Volcanism and geochemistry in Central America: progress and prob- ciated with flow of asthenosphere beneath southern Central America. Geology
lems. In: Carr, M.J., Feigenson, M.D., Patino, L.C., Eiler, J. (Eds.), Inside the Subduction 23 (7):617–620. http://dx.doi.org/10.1130/0091-7613(1995)023b0617:
Factory. 2004. American Geophysical Union, Washington, D. C:pp. 153–174. http:// VILCAWN2.3.CO;2.
dx.doi.org/10.1029/138GM09. Hidalgo, H., Alfaro, E., Quesada-Montano, B., 2017. Observed (1970–1999) climate vari-
Carr, M.J., Saginor, I., Alvarado, G.E., Bolge, L.L., Lindsay, F.N., Milidakis, K., Turrin, B.D., ability in Central America using a high-resolution meteorological dataset with impli-
Feigenson, M.D., Swisher III, C.C., 2007. Element fluxes from the volcanic front of Nic- cation to climate change studies. Clim. Chang. 141 (13). http://dx.doi.org/10.1007/
aragua and Costa Rica. Geochem. Geophys. Geosyst. 8 (Q06001):1–22. http:// s10584-016-1786-y.
dx.doi.org/10.1029/2006GC001396. Hoernle, K., Abt, D.L., Fischer, K.M., Nichols, H., Hauff, F., Abers, G.A., Van Den Bogaard, P.,
Chiodini, G., Marini, L., Russo, M., 2001. Geochemical evidence for the existence of high- Heydolph, K., Alvarado, G., Protti, M., Strauch, W., 2008. Arc-parallel flow in the man-
temperature hydrothermal brines at Vesuvio volcano, Italy. Geochim. Cosmochim. tle wedge beneath Costa Rica and Nicaragua. Nature 451 (7182):1094–1097. http://
Acta 65 (13):2129–2147. http://dx.doi.org/10.1016/S0016-7037(01)00583-X. dx.doi.org/10.1038/nature06550.
Craig, H., 1961. Isotopic variations in meteoric waters. Science 133 (3465):1702–1703. Husen, S., Quintero, R., Kissling, E., Hacker, B., 2003. Subduction-zone structure and mag-
http://dx.doi.org/10.1126/science.133.3465.1702. matic processes beneath Costa Rica constrained by local earthquake tomography and
Craig, H., 1963. The isotopic geochemistry of water and carbon in geothermal areas. Nucle- petrological modelling. Geophys. J. Int. 155 (1):11–32. http://dx.doi.org/10.1046/
ar Geology on Geothermal Areas, Spoleto Conference Proceedings. 1963, pp. 17–53. j.1365-246X.2003.01984.x.
Criss, R.E., 1999. Principles of Stable Isotope Distribution. Oxford University Press. Instituto Nicaragüense de Estudios Territoriales (INETER), 2005. Clima de Nicaragua.
Dansgaard, W., 1964. Stable isotopes in precipitation. Tellus 16 (4):436–468. http:// Available at:. http://www.ineter.gob.ni/met.html, Accessed date: 7 June 2017.
dx.doi.org/10.3402/tellusa.v16i4.8993. Instituto Nicaragüense de Estudios Territoriales (INETER), 2017. Clima de Nicaragua.
Delcamp, A., Roberti, G., van Wyk de Vries, B., 2016. Water in volcanoes: evolution, stor- Available at:. http://www.ineter.gob.ni/met.html, Accessed date: 7 June 2017.
age and rapid release during landslides. Bull. Volcanol. 78 (87). http://dx.doi.org/ International Atomic Energy Agency (IAEA), 1992. Estudios geotérmicos con técnicas
10.1007/s00445-016-1082-8. isotópicas y geoquímicas en América Latina. IAEA-TECDOC-641. Vienna, Austria
DeMets, C., 2001. A new estimate for present-day Cocos-Caribbean plate motion: implica- (452 pp). http://www.iaea.org/inis/collection/NCLCollectionStore/_Public/23/047/
tions for slip along the Central American volcanic arc. Geophys. Res. Lett. 28 (21): 23047525.pdf.
4043–4046. http://dx.doi.org/10.1029/2001GL013518. Kruskal, W.H., Wallis, W.A., 1952. Use of ranks in one-criterion variance analysis. J. Am.
Denyer, P., Gazel, E., 2009. The Costa Rican Jurassic to Miocene oceanic complexes: origin, Stat. Assoc. 47 (260):583–621. http://www.jstor.org/stable/2280779.
tectonics and relations. J. S. Am. Earth Sci. 28 (4):429–442. http://dx.doi.org/10.1016/ Landwehr, J.M., Coplen, T.B., 2006. Line-conditioned excess: a new method for character-
j.jsames.2009.04.010. izing stable hydrogen and oxygen isotope ratios in hydrologic systems. Isotopes in
Dunn, O.J., 1961. Multiple comparisons among means. J. Am. Stat. Assoc. 56 (293), 52–64. Environmental Studies :pp. 132–135. https://inis.iaea.org/search/search.aspx?orig_
Elkins, L.J., Fischer, T.P., Hilton, D.R., Sharp, Z.D., McKnight, S., Walker, J., 2006. Tracing nitro- q=RN:37043527.
gen in volcanic and geothermal volatiles from the Nicaraguan volcanic front. Geochim. Leeman, W.P., Carr, M.J., Morris, J.D., 1994. Boron geochemistry of the Central Amer-
Cosmochim. Acta 70 (20):5215–5235. http://dx.doi.org/10.1016/j.gca.2006.07.024. ican Volcanic Arc: constraints on the genesis of subduction-related magmas.
Esquivel, G., Sánchez, R., Birkel, C., Good, S.P., Boll, J., 2017. Hydroclimatic and Geochim. Cosmochim. Acta 58 (1):149–168. http://dx.doi.org/10.1016/0016-
ecohydrological resistance/resilience conditions across tropical biomes of Costa 7037(94)90453-7.
Rica. Ecohydrology http://dx.doi.org/10.1002/eco.1860. LeGrande, A.N., Schmidt, G.A., 2006. Global gridded data set of the oxygen isotopic com-
Fehn, U., Snyder, G.T., Varekamp, J.C., 2002. Detection of recycled marine sediment com- position in seawater. Geophys. Res. Lett. 33, L12604. http://dx.doi.org/10.1029/
ponents in crater lake fluids using 129 I. J. Volcanol. Geotherm. Res. 115 (3):451–460. 2006GL026011.
http://dx.doi.org/10.1016/S0377-0273(01)00325-0. López, D.L., Bundschuh, J., Soto, G.J., Fernandez, J.F., Alvarado, G.E., 2006. Chemical evolu-
Feigenson, M.D., Carr, M.J., 1993. The source of Central American lavas: inferences from tion of thermal springs at Arenal Volcano, Costa Rica: effect of volcanic activity, pre-
geochemical inverse modeling. Contrib. Mineral. Petrol 113:226–235. http:// cipitation, seismic activity, and Earth tides. J. Volcanol. Geotherm. Res. 157 (1):
dx.doi.org/10.1007/BF00283230. 166–181. http://dx.doi.org/10.1016/j.jvolgeores.2006.03.049.
Feigenson, M.D., Carr, M.J., Maharaj, S.V., Juliano, S., Bolge, L.L., 2004. Lead isotope compo- Lucic, G., Stix, J., Lollar, B.S., Lacrampe-Couloume, G., Muñoz, A., Carcache, M.I., 2014. The
sition of Central American volcanoes: influence of the Galapagos plume. Geochem. degassing character of a young volcanic center: Cerro Negro, Nicaragua. Bull.
Geophys. Geosyst. 5 (6). http://dx.doi.org/10.1029/2003GC000621. Volcanol. 76 (9):850. http://dx.doi.org/10.1007/s00445-014-0850-6.
Fischer, T.P., Hilton, D.R., Zimmer, M.M., Shaw, A.M., Sharp, Z.D., Walker, J.A., 2002. Sub- Madrigal-Solís, H., Fonseca-Sánchez, A., Reynolds-Vargas, J., 2017. Hydrogeochemical
duction and recycling of nitrogen along the Central American margin. Science 297 characterization of Barva and Colima aquifers in the Central Valley of Costa Rica.
(5584):1154–1157. http://dx.doi.org/10.1126/science.1073995. Water Technol. Sci. (in Spanish) 8 (1), 115–132.
Fournier, R.O., Hanshaw, B.B., Urrutia Sole, J.F., 1982. Oxygen and hydrogen isotopes in thermal Magaña, V., Amador, J.A., Medina, S., 1999. The midsummer drought over Mexico and
waters at Zunil, Guatemala. Trans.-Geotherm. Resour. Counc., United States, pp. 89–91. Central America. Climate 12:1577–1588. http://dx.doi.org/10.1175/1520-
Fridleifsson, I.B., 2001. Geothermal energy for the benefit of the people. Renew. Sust. 0442(1999)012b1577:TMDOMAN2.0.CO;2.
Energ. Rev. 5 (3):299–312. http://dx.doi.org/10.1016/S1364-0321(01)00002-8. Maldonado, T., Alfaro, E., Fallas, B., Alvarado, L., 2013. Seasonal prediction of extreme pre-
Gazel, E., Carr, M.J., Hoernle, K., Feigenson, M.D., Szymanski, D., Hauff, F., van den Bogaard, cipitation events and frequency of rainy days over Costa Rica, Central America, using
P., 2009. Galapagos-OIB signature in southern Central America: mantle refertilization Canonical Correlation Analysis. Adv. Geosci. 33 (33):41–52. http://dx.doi.org/
by arc–hot spot interaction. Geochem. Geophys. Geosyst. 10 (2), Q02S11. http:// 10.5194/adgeo-33-41-2013.
dx.doi.org/10.1029/2008GC002246. Manzo, A.R.R., 2005. Geothermal power development in Guatemala 2000–2005. Proceed-
Gazel, E., Hoernle, K., Carr, M.J., Herzberg, C., Saginor, I., van den Bogaard, P., Hauff, F., ings World Geothermal Congress 2005 Antalya, Turkey, 24–29 April 2005.
Feigenson, M.D., Swisher III, C., 2011. Arc-plume interaction in Central America: In- Marini, L., Fung, A.Y., Sanchez, E., 2003. Use of reaction path modeling to identify the
flux of Galapagos asthenosphere and slab melting. Lithos 121:117–134. http:// processes governing the generation of neutral Na–Cl and acidic Na–Cl–SO 4 deep
dx.doi.org/10.1016/j.lithos.2010.10.008. geothermal liquids at Miravalles geothermal system, Costa Rica. J. Volcanol.
Gazel, E., Hayes, J.L., Hoernle, K., Kelemen, P., Everson, E., Holbrook, W.S., Hauff, F., van den Geotherm. Res. 128 (4):363–387. http://dx.doi.org/10.1016/S0377-
Bogaard, P., Vance, E.A., Chu, S., Calvert, A.J., Carr, M.J., Yogodzinski, G.M., 2015. Con- 0273(03)00226-9.
tinental crust generated in oceanic arcs. Nat. Geosci. 8 (4):321–327. http://dx.doi.org/ Martínez-Tiffer, E.M., Lacayo, R.A., Sabatino, G., 1988. Geothermal development in
10.1038/ngeo2392. Nicaragua. Geothermics 17 (2–3):333–354. http://dx.doi.org/10.1016/0375-
Giggenbach, W.F., 1992. Isotopic shifts in waters from geothermal and volcanic systems 6505(88)90062-4.
along convergent plate boundaries and their origin. Earth Planet. Sci. Lett. 113 (4): Molina, F., Martí, J., 2016. The Borinquen geothermal system (Cañas Dulces caldera, Costa Rica).
495–510. http://dx.doi.org/10.1016/0012-821X(92)90127-H. Geothermics 64:410–425. http://dx.doi.org/10.1016/j.geothermics.2016.07.001.
A. Ramírez-Leiva et al. / Journal of Volcanology and Geothermal Research 345 (2017) 172–183 183

Moya, P., Sánchez, E., América, C., 2005. Non-condensable gases at the Miravalles geother- Saginor, I., Gazel, E., Condie, C., Carr, M.J., 2013. Evolution of geochemical variations along
mal field. Proceedings of the Thirtieth Workshop on Geothermal Reservoir Engineer- the Central American volcanic front. Geochem. Geophys. Geosyst. 14 (10):
ing https://pangea.stanford.edu/ERE/pdf/IGAstandard/SGW/2005/moya2.pdf. 4504–4522. http://dx.doi.org/10.1002/ggge.20259.
Nieva, D., Verma, M.P., Santoyo, E., Portugal, E., Campos, A., 1997. Geochemical explora- Sak, P.B., Fisher, D.M., Gardner, T.W., Marshall, J.S., LaFemina, P.C., 2009. Rough crust sub-
tion of the Chipilapa geothermal field, El Salvador. Geothermics 26 (5–6):589–612. duction, forearc kinematics, and Quaternary uplift rates, Costa Rican segment of the
http://dx.doi.org/10.1016/S0375-6505(97)00012-6. Middle American Trench. Geol. Soc. Am. Bull. 121 (7–8):992–1012. http://
Peacock, S.M., van Keken, P.E., Holloway, S.D., Hacker, B.R., Abers, G.A., Fergason, R.L., dx.doi.org/10.1130/B26237.1.
2005. Thermal structure of the Costa Rica–Nicaragua subduction zone. Phys. Earth Sánchez-Murillo, R., Birkel, C., 2016. Groundwater recharge mechanisms inferred from
Planet. Inter. 149 (1):187–200. http://dx.doi.org/10.1016/j.pepi.2004.08.030. isoscapes in a complex tropical mountainous region. Geophys. Res. Lett. 43 (10):
Pearson, C.P., Connor, C., Sanford, W.E., 2008. Rapid response of a hydrologic system to 5060–5069. http://dx.doi.org/10.1002/2016GL068888.
volcanic activity: Masaya volcano, Nicaragua. Geology 36 (12):951–954. http:// Sánchez-Murillo, R., Esquivel, G., Welsh, K., Brooks, E., Boll, J., Alfaro, R., Valdés, J., 2013.
dx.doi.org/10.1130/G25210A.1. Spatial and temporal variation of stable isotopes in precipitation across Costa Rica:
Phillips-Lander, C.M., Fowle, D.A., Taunton, A., Hernandez, W., Mora, M., Moore, D., an analysis of historic GNIP records. J. Mod. Hydrol. 3:226–240. http://dx.doi.org/
Shinogle, H., Roberts, J.A., 2014. Silicate dissolution in Las Pailas thermal field: impli- 10.4236/ojmh.2013.34027.
cations for microbial weathering in acidic volcanic hydrothermal spring systems. Sánchez-Murillo, R., Esquivel-Hernández, G., Sáenz-Rosales, O., Piedra-Marín, G., Fonseca-
Geomicrobiol J. 31 (1):23–41. http://dx.doi.org/10.1080/01490451.2013.802395. Sánchez, A., Madrigal-Solís, H., Ulloa-Chaverri, F., Rojas-Jiménez, L.D., Vargas-Víquez,
Protti, M., Guendel, F., McNally, K., 1995. Correlation between the age of the subducting J.A., 2016. Isotopic composition in precipitation and groundwater in the northern
Cocos Plate and the geometry of the Wadati-Benioff Zone under Nicaragua and Costa mountainous region of the Central Valley of Costa Rica. Isot. Environ. Health Stud.
Rica. Geol. Soc. Am. Spec. Pap. 295:309–327. http://dx.doi.org/10.1130/SPE295-p309. http://dx.doi.org/10.1080/10256016.2016.1193503.
Protti, M., González, V., Freymueller, J., Doelger, S., 2012. Isla del Coco, on Cocos Plate, con- Shaw, A.M., Hilton, D.R., Fischer, T.P., Walker, J.A., Alvarado, G.E., 2003. Contrasting He–C
verges with Isla de San Andrés, on the Caribbean Plate, at 78mm/yr. Int. J. Trop. Biol. relationships in Nicaragua and Costa Rica: insights into C cycling through subduction
60, 33–41. zones. Earth Planet. Sci. Lett. 214 (3):499–513. http://dx.doi.org/10.1016/S0012-
R Core Team, 2014. R: A language and environment for statistical computing. R Foundation 821X(03)00401-1.
for Statistical Computing, Vienna, Austria Available at:. http://www.R-project.org/. Simkin, T., Siebert, L., Kimberly, P., 2011. Volcanoes of the World. Third Edition. University
Ranero, C.R., von Huene, R., 2000. Subduction erosion along the Middle America conver- of California Press and Smithsonian Institute, USA, p. 479.
gent margin. Nature 404:748–752. http://dx.doi.org/10.1038/35008046. Solé, J.M., Arasa, R., Picanyol, M., González, M.Á., Domingo-Dalmau, A., Masdeu, M., Porras,
Reagan, M.K., Gill, J.B., 1989. Coexisting of calc-alkaline and high-Nb basalts from Turrialba I., Codina, B., 2016. Assessment of climate change in Nicaragua: analysis of precipita-
volcano, Costa Rica: implications for residual titanates in arc magma sources. tion and temperature by dynamical downscaling over a 30-year horizon. Atmos. Cli-
J. Geophys. Res. Solid Earth Planets 94:4619–4633. http://dx.doi.org/10.1029/ mate Sci. 6:445–474. http://dx.doi.org/10.4236/acs.2016.63036.
JB094iB04p04619. Sprenger, M., Tetzlaff, D., Tunaley, C., Dick, J., Soulsby, C., 2017. Evaporation fractionation
Rizzo, A.L., Di Piazza, A., de Moor, J.M., Alvarado, G.E., Avard, G., Carapezza, M.L., Mora, M.M., in a peatland drainage network affects stream water isotope composition. Water
2016. Eruptive activity at Turrialba volcano (Costa Rica): inferences from 3He/4He in Resour. Res. 53:851–866. http://dx.doi.org/10.1002/2016WR019258.
fumarole gases and chemistry of the products ejected during 2014 and 2015. Geochem. Taran, Y.A., Pokrovsky, B.G., Esikov, A.D., 1989. Deuterium and oxygen-18 in fumarolic
Geophys. Geosyst. 17:4478–4494. http://dx.doi.org/10.1002/2016GC006525. steam and amphiboles from some Kamchatka volcanoes: “andesitic waters”. Dokl.
Rouwet, D., Inguaggiato, S., Taran, Y., Varley, N., 2009. Chemical and isotopic compositions Akad. Nauk SSSR. 304, pp. 440–443.
of thermal springs, fumaroles and bubbling gases at Tacaná Volcano (Mexico–Guate- Tassi, F., Vaselli, O., Capaccioni, B., Giolito, C., Duarte, E., Fernandez, E., Minissale, A., Magro,
mala): implications for volcanic surveillance. Bull. Volcanol. 71 (3):319. http:// G., 2005. The hydrothermal-volcanic system of Rincon de la Vieja volcano (Costa
dx.doi.org/10.1007/s00445-008-0226-x. Rica): a combined (inorganic and organic) geochemical approach to understanding
Rouwet, D., Mora, R., Ramírez, C.J., González, g. and Inguagguati, D., 2016. Dynamic fluid the origin of the fluid discharges and its possible application to volcanic surveillance.
recycling at Laguna Caliente (Poás, Costa Rica) before and during the 2006 ongoing J. Volcanol. Geotherm. Res. 148 (3):315–333. http://dx.doi.org/10.1016/
phreatic eruption cycle (2005-10). Geol. Soc. 473 (1):73–96. http://dx.doi.org/ j.jvolgeores.2005.05.001.
10.1144/SP437.11. von Huene, R., Scholl, D.W., 1991. Observations at convergent margins concerning sedi-
Rowe, G.L., Brantley, S.L., Fernandez, J.F., Borgia, A., 1995. The chemical and hydrologic ment subduction, subduction erosion, and the growth of continental crust. Rev.
structure of Poa's Volcano, Costa Rica. J. Volcanol. Geotherm. Res. 64 (3):233–267. Geophys. 29 (3):279–316. http://dx.doi.org/10.1029/91RG00969.
http://dx.doi.org/10.1016/0377-0273(94)00079-V. von Huene, R., Ranero, C.R., Weinrebe, W., Hinz, K., 2000. Quaternary convergent margin
Ryan, Jeffrey G., Chauvel, C., 2013. The subduction-zone filter and the impact of recycled tectonics of Costa Rica, segmentation of the Cocos plate, and Central American volca-
materials on the evolution of the mantle. Treatise on Geochemistry (Second Edition). nism. Tectonics 19:314–334. http://dx.doi.org/10.1029/1999TC001143.
3. School of Geosciences Faculty and Staff Publications. 888:pp. 479–508. http:// Waylen, P.R., Caviedes, C.N., Quesada, M., 1996. Interannual variability of monthly precip-
dx.doi.org/10.1016/B978-0-08-095975-7.00211-4. itation in Costa Rica. Climate 9 (10):2606–2613. http://dx.doi.org/10.1175/1520-
Sáenz, F., Durán, A.M., 2015. A climatology of low level wind regimes over Central Amer- 0442(1996)009b2606:IVOMPIN2.0.CO;2.
ica using weather type classification approach. Front. Earth Sci. 3 (15):1–18. http:// Zamboni, D., Gazel, E., Ryan, J.G., Cannatelli, C., Lucchi, F., Atlas, Z.D., Trela, J., Mazza, S.E.,
dx.doi.org/10.3389/feart.2015.00015. De Vivo, B., 2016. Contrasting sediment melt and fluid signatures for
Saginor, I., 2008. Volcanic history of western Nicaragua and geochemical evolution of the magmacomponents in the Aeolian Arc: implications for numerical modeling of sub-
Central American volcanic front. Rutgers. The State University of New Jersey-New duction systems. Geochem. Geophys. Geosyst. 17:2034–2053. http://dx.doi.org/
Brunswick. 10.1002/2016GC006301.
Saginor, I., Gazel, E., Carr, M.J., Swisher III, C.C., Turrin, B., 2011. New Pliocene-Pleistocene Zimmer, M.M., Fischer, T.P., Hilton, D.R., Alvarado, G.E., Sharp, Z.D., Walker, J.A., 2004. Ni-
(40)Ar/(39)Ar ages fill in temporal gaps in the Nicaraguan volcanic record. trogen systematics and gas fluxes of subduction zones: insights from Costa Rica arc
J. Volcanol. Geotherm. Res. 202 (1–2):143–152. http://dx.doi.org/10.1016/ volatiles. Geochem. Geophys. Geosyst. 5 (Q05J11). http://dx.doi.org/10.1029/
j.jvolgeores.2011.02.002. 2003GC000651.

You might also like